Skip to main content

Using the collaborative cross to identify the role of host genetics in defining the murine gut microbiome

Abstract

Background

The human gut microbiota is a complex community comprised of trillions of bacteria and is critical for the digestion and absorption of nutrients. Bacterial communities of the intestinal microbiota influence the development of several conditions and diseases. We studied the effect of host genetics on gut microbial composition using Collaborative Cross (CC) mice. CC mice are a panel of mice that are genetically diverse across strains, but genetically identical within a given strain allowing repetition and deeper analysis than is possible with other collections of genetically diverse mice.

Results

16S rRNA from the feces of 167 mice from 28 different CC strains was sequenced and analyzed using the Qiime2 pipeline. We observed a large variance in the bacterial composition across CC strains starting at the phylum level. Using bacterial composition data, we identified 17 significant Quantitative Trait Loci (QTL) linked to 14 genera on 9 different mouse chromosomes. Genes within these intervals were analyzed for significant association with pathways and the previously known human GWAS database using Enrichr analysis and Genecards database. Multiple host genes involved in obesity, glucose homeostasis, immunity, neurological diseases, and many other protein-coding genes located in these regions may play roles in determining the composition of the gut microbiota. A subset of these CC mice was infected with Salmonella Typhimurium. Using infection outcome data, an increase in abundance of genus Lachnospiraceae and decrease in genus Parasutterella correlated with positive health outcomes after infection. Machine learning classifiers accurately predicted the CC strain and the infection outcome using pre-infection bacterial composition data from the feces.

Conclusion

Our study supports the hypothesis that multiple host genes influence the gut microbiome composition and homeostasis, and that certain organisms may influence health outcomes after S. Typhimurium infection.

Video Abstract

Introduction

The nutrient-rich environment of the human intestinal tract harbors up to 100 trillion microbes [1]. Sterile at birth, our colon becomes densely populated with microbes, in the range of 1011–1012 cells/ml, the highest recorded density for any microbial habitat [2]. These microbes become an integral part of the digestive system breaking down complex molecules, modifying host-derived molecules (like bile acids for example), and modulating the immune response [3].

Differential abundance of certain microbes in the intestine has been implicated in various diseases, including inflammatory bowel disease (IBD), obesity, and food allergies [4,5,6]. Some microbial metabolites interact with the gut-brain axis and have been associated with autism [7], Parkinson’s disease [8], Alzheimer’s disease [9], epilepsy [10], and stroke [11]. Other microbes influence circadian rhythm disturbances, depression, and disruption of sleep patterns [12]. Despite these associations, establishing causal relationships between a microbiome and a disease is challenging.

The recombinant inbred panel of mice called the Collaborative Cross (CC) was created to model a genetically diverse population. Eight founder strains, A/J, C57BL/6 J, 129S1/SvImJ, NOD/LtJ, NZO/HiLtJ, CAST/Ei, PWK/PhJ, and WSB/EiJ, were bred in a funnel fashion to generate a large panel of inbred strains with balanced allele frequency and dense, evenly distributed recombination sites. The CC panel approximates the genetic diversity found in humans and has a high resolution for mapping quantitative trait loci (QTL) [13,14,15]. CC mice have been widely used to define host genetic loci implicated in various infectious diseases including West Nile virus, tuberculosis, and influenza as well as other phenotypes including glucose tolerance, DNA damage, and Epilepsy [16,17,18,19,20,21,22]. In Ebola infections in CC mice, a diverse phenotypic variation ranging from high resistance to complete lethality was observed across CC strains, and a central transcriptional regulatory gene called Tek (TEK receptor tyrosine kinase) correlated with weight loss and mortality after infection [17]. In CC mice, liver and spleen bacterial load varies across strains after infection with S. Typhimurium (STm) [23]. A candidate gene Slc35f1 (solute carrier family 35, member F1), which has lactate dehydrogenase activity and is predicted to influence the pyruvate metabolism pathway in STm, was identified as potentially important [24].

Though the microbiome has traditionally been seen as influenced by the environment, recent studies have linked the influence of host genetics to microbiome composition [25]. In human twin studies, bacterial representation was more similar in monozygotic twins than in dizygotic twins, supporting a positive association between genetic factors and microbiome [26]. Re-analysis of data from previous human twin studies has established a similar association with the host genetics and the microbiome [26,27,28]. The effects of genetics, diet, and the environment make establishing causal relationships challenging in human studies.

In murine models with controlled diet and environment, a relationship between the host genetic makeup and the microbiome is becoming apparent. QTL analysis on the advanced intercross mouse population has shown that microbial abundance is a polygenic trait [29, 30]. Pleiotropy in QTL regions was observed for both closely related and unrelated bacteria. Other mouse population analyses identified several microbial abundance QTLs and correlated them to immune response, obesity, and insulin secretion genes from the host [31, 32]. The CC population has also been used to associate microbial metabolites with GI cancers, lipid metabolism, and inflammation [33]. Recently, the intestinal microbiome composition of CC mice has been associated with a sleep phenotype, memory, anxiety-like behavior, and Azoxymethane-induced toxicity [34,35,36,37].

Over a 2-year time period, we collected feces from 28 CC mouse strains, including three males and three females of each strain, and performed 16s rRNA gene sequencing. Individuals within a strain had similar microbiota composition, while there were significant differences between the strains starting at the phylum level. Microbial abundance data at the genus level identified significant genes associated with individual genera across the genome. After collection of feces, a subset of CC strains was infected with Salmonella enterica serotype Typhimurium (STm) in a parallel series of experiments [23]. We used the survival data after STm infection to identify and correlate the microbes associated with STm infection outcome. A machine learning algorithm predicted the correct CC strain and the infection outcome using bacterial composition data. The increase in abundance of genus Lachnospiraceae and decrease in genus Parasutterella correlated with a positive health outcome after infection.

Methods

Animals

Eight- to 12-week-old male and female Collaborative Cross mice (CC) were bred at the Division of Comparative Medicine at Texas A&M University. In preparation for a different series of experiments, mice were implanted with an E-mitter telemetry device (STARR Life Sciences Corp.) and permitted to recover for 1 week (detailed methodology described in [23]). All mice were transferred to a second facility and individually housed at least 5 days before the collection of feces. Individually housed mice had hardwood chip bedding in ventilated cages, with nestlet squares, a cardboard hut, and were fed a standardized rodent diet (Teklad Global 19% protein extruded rodent diet, irradiated, 2919, from ENVIGO) and sterile water ad libitum.

Ethics statement

Mouse studies were conducted in accordance with the Guide for the Care and Use of Laboratory Animals of the National Institutes of Health. All mouse studies were conducted at the Texas A&M Health Science Center using protocols (AUP—2015–0315D and 2018–0488D) approved by the Texas A&M Institutional Animal Care and Use Committee (IACUC).

Fecal collection, storage, and DNA extraction

Feces were collected by placing mice on a paper towel under an inverted glass beaker. Sterile forceps were used to transfer at least three fecal pellets per animal to cryovials and feces were stored at − 80 °C until use. Total DNA from one fecal pellet for each animal was extracted using MagAttract PowerSoil DNA EP Kit from Qiagen.

Salmonella infection

CC mice were infected by gavage with virulent wild-type Salmonella enterica serotype Typhimurium strain HA420 (ATCC14028s nalr) as previously described [23, 38]. Briefly, HA420 was grown to the stationary phase at 37 °C with aeration, serially diluted, and plated for bacterial CFU to determine the exact titer of the inoculum. One week after moving to a different facility, mice were inoculated by gavage with approximately 2–5 × 107 CFU of S. Typhimurium HA420 in 100 µl of LB broth. Body temperature and activity data from the telemetry device, body condition scores, and weight loss data from daily health checks were used to monitor disease progression. Mice that developed severe clinical symptoms were humanely euthanized by CO2 asphyxiation. If mice remained healthy at the end of the experiment, they were euthanized at 7 days post-infection.

Sequencing of 16S rRNA genes

Amplicon libraries were prepared for the variable V3 and V4 regions of the prokaryotic 16S rRNA gene. Briefly, extracted microbial DNA from one fecal pellet per animal was amplified using 16S amplicon PCR forward primer = 5′ TCGTCGGCAGCGTCAGATGTGTATAAGAGACAGCCT.

ACGGGNGGCWGCAG and 16S amplicon PCR reverse primer = 5′ GTCTCGTGGGCTCGGA.

GATGTGTATAAGAGACAGGACTACHVGGGTATCTAATCC in a BIORAD thermocycler using the following cycling conditions (denaturation at 95 °C for 3 min followed by 25 cycles of 95 °C for 30 s, 55 °C for 30 s, 72 °C for 30 s, and a final extension at 72 °C for 5 min). The PCR products were purified with AMPure XP beads to purify the V3 and V4 amplicon to be free from primers and primer-dimer species. The purified amplicons were dual indexed using a Nextera XT Index kit and sequenced using the Illumina Miseq platform at the Texas A&M Institute of Genome Sciences and Society (TIGSS) to generate 2 x300 base pair (bp) paired-end sequences.

Preprocessing of sequences

A DADA2 pipeline was used for processing the 16S Illumina amplicon sequence data. Forward reads were trimmed at 17 and 285 bp and reserve reads were trimmed at 21 and 205 bp to remove the primers and bases with a median quality score of less than 25. Denoising, merging, and removal of chimeric sequences reduced the total sequences from 31,405,204 to 11,346,536 (Fig S1). After removing the Amplicon Sequence Variants (ASVs) that were not present in at least two samples, and mitochondrial and chloroplast sequences, 1637 unique ASVs were obtained.

Data processing

Metagenomic bioinformatic analysis was performed using QIIME 2 2020.6 [39]. Raw sequence data were demultiplexed and quality filtered using the q2-demux plugin followed by denoising with DADA2 (via q2-dada2) [40]. Taxonomy was assigned to Amplicon Sequence Variants (ASVs) using the q2-feature-classifier [41] classify-sklearn naïve Bayes taxonomy classifier against Silva 138 SSURef NR99 full-length OTU reference sequences specifically trained using the 16S rRNA gene sequencing primers at 99% identity [42]. The feature table was filtered to include only the assigned reads of the kingdom bacteria and to remove singleton features and mitochondrial and chloroplast sequences. All amplicon sequence variants (ASVs) were aligned with mafft [43] (via q2-alignment) and used to construct a phylogeny with fasttree2 [44] (via q2-phylogeny). Samples were rarefied (subsampled without replacement) to 21,250 sequences per sample to reduce the effect of sampling depth. Alpha-diversity metric (Shannon’s Index), beta diversity metrics (unweighted UniFrac [45], and Bray–Curtis dissimilarity [46]), and principal coordinate analysis (PCoA) were estimated using q2-diversity. Group significance between alpha and beta diversity indexes were calculated with QIIME2 plugins using the Kruskal–Wallis test and permutational multivariate analysis of variance (PERMANOVA), respectively [47]. Additionally, analysis of similarity (ANOSIM) was performed on the beta diversity values [48]. Specific cage information for these animals was lost during a system shift in our animal facility. Therefore, we used the date of animal arrival to our facility over a 2-year period, combined with strain to calculate littermate effects.

ANCOM

16S analysis of a single hypervariable region does not provide accurate classification at the species level. Thus, we used genus-level data. The abundance table was reduced to genus level and taxa that were not present in at least four different CC strains were removed. This table was used for further downstream analysis. Differential abundance testing was performed using ANCOM, to compare the relative abundance of a taxon between two groups by performing statistical tests on data transformed by an additive log-ratio of the abundance of a given taxon versus the abundance of all other taxa individually [49]. The analysis was performed using the R package—ANCOMBC, considering structural zeros grouped by strain and sex as covariates [50].

Heritability

Broad-sense heritability was calculated using the formula H2 = VG/VP = VG/(VE + VG) as previously described [51, 52]. VG and VE are the variance explained by the genetic and the environmental component respectively while VP is the total phenotypic variance for a given trait. Genus-level abundance data was used for this analysis. For each phenotype, VG was calculated as the mean variance between the replicates of the same strain. VP was calculated as the variance across all the strains for each corresponding phenotype. The maximum value for heritability is one.

QTL analysis

QTL analysis was performed using R/qtl2 software [53]. This method performs genome scans using a linear mixed model to account for the complex population structure in the CC mice. Rank-transformed genus-level abundance was used as the phenotype and the genotypes were imputed from QTLViewer [54]. A genome-wide scan was performed using the scan1 function. The generated Logarithm of Odds (LOD) score is the likelihood ratio comparing the hypothesis of a QTL at a given position versus that of no QTL. To establish genome-wide significance, each individual phenotype was randomly shuffled 999 times and LOD scores were calculated for each iteration. Values at or above the 85th percentile were considered significant for that phenotype.

Candidate gene selection and enrichment

The Genomic confidence interval was calculated by dropping the LOD scores by 1.8, for each significant peak. The Mouse Genome Informatics (MGI) was used to find genes and QTL features within each interval [55]. To further shortlist candidate genes, the founder strain distribution pattern was queried against the CC variant database (V3 version) [56]. Variants were further shortlisted based on the impact scores calculated by the Ensembl Variant Effect Predictor (VEP) [57]. GeneCards database was used to collect human gene summaries, previously known human GWAS and Super pathways [58]. Gene enrichment analysis was performed using Enrichr [59, 60]. Briefly, genes with variants that matched the founder strain distribution pattern were imputed in the Enrichr web interface. Data tables for KEGG Human pathways 2021 and MGI mammalian phenotype 2021 were downloaded. Terms with a P-value of < 0.05 were considered significant.

Machine learning

A random forest classifier (scikit-learn 0.24.2) was used to perform the machine learning analysis on the genus-level data [61]. For the strain classification, one individual per strain was randomly selected as test data and the classifier was trained on the remaining data. The test data was then used to predict the accuracy. Cross-validation was performed using all the members of the training data. The accuracy of the model is the average score of the cross-validated model. For health classification, a similar model was used with 25% test data and 75% training data. The area under the receiver operating characteristics (AUROC), a probability curve based on the degree of separability, was also used as a performance measurement for this trained classifier along with average cross-validated accuracy.

Results

Composition of the bacterial community is variable across CC strains

One hundred sixty seven mice aged 8–12 weeks, 3 males and 3 females representing 28 CC strains (except for CC078—only two females), were included in this study. We noted a large variability in the bacterial composition of the feces across the CC strains that we studied (Fig. 1A). After assigning taxonomy using the SILVA 138 reference database, the Naïve Bayes classifier identified 10 phyla, 14 classes, 20 orders, 47 families, 113 genera, and 236 species with 99% sequence identity. At level 2, Bacteroidota (60.96%) was the most abundant phylum, followed by Firmicutes (34.52%) and Pastescibacteria (1.29%). The remaining seven phyla constituted less than 4% of the overall microbial composition and presence varied across strains (Table 1). Muribaculaceae, followed by Lachnospiraceae and Alistipes, were the top three genera by total counts. Only 20 genera were present in at least 95% of the samples and 37 genera were present in less than 25% of the samples (Table S1). This high variability both at the phylum and the genus level reflects the diversity of gut microbiome across the CC strains.

Fig. 1
figure 1

Diversity across the CC strain starts at the phylum Level (L2). A Bar graph of the relative bacterial abundance of all the phyla present in feces of 28 CC strains. Data represent 3 males and 3 females of each strain. B Within sample diversity measured by Shannon index. Bars represent mean and standard deviation for each strain. Kruskal–Wallis test was performed to analyze the statistical significance (p < 0.001)

Table 1 Bacteroidota and Firmicutes were the most abundant phylum across the CC strains. Average, minimum, and maximum abundance of various phyla by strain, arranged in order through decreasing overall abundance

Bacterial communities within a strain are more similar than across strains

The Shannon index, a diversity measure based on richness, was calculated for individual animals (Fig. 1B). Individuals within a strain had similar diversity values while the values varied widely across the strains. The average diversity values for strains ranged from 5.7 to 7 with a mean value of 6.2 and a standard deviation of 0.3. The range of these values further demonstrates the diversity across the CC strains. Pairwise Kruskal–Wallis tests showed a significant difference (P < 0.001) in Shannon index across CC strains.

A non-phylogenetic quantitative metric, Bray–Curtis dissimilarity, and phylogenetic qualitative metric, unweighted UniFrac, were used to measure beta diversity. Principal component analysis (PCOA) plots of Bray–Curtis dissimilarity and the unweighted UniFrac distance metric groups animals better within a strain than any other parameter in the metadata (Fig. 2A, B). To test for significance, pairwise PERMANOVA and ANOSIM analyses were performed. For Bray–Curtis, both analyses showed significant differences between the CC strains using the Kruskal–Wallis test. Unweighted UniFrac showed similar significance, but lower R and F values compared to Bray–Curtis (Table 2). This decrease could be explained by the complete absence of several ASV’s in some of the CC strains, as unweighted uniFrac takes presence/absence data into account.

Fig. 2
figure 2

Analysis of beta diversity suggests fecal microbiota are more similar within CC strains than across CC strains. Principal component analysis (PCoA) 2D plots of between-sample dissimilarities measured by Bray–Curtis distance (A) and unweighted UniFrac distance (B). Each color represents a CC strain. Box plot of distances between fecal microbial communities obtained when comparing mice within and between strains of CC mice using Bray–Curtis distance (C) and Unweighted UniFrac distance (D). P < 0.05(*) for Wilcoxon signed rank test

Table 2 Bacterial composition differs across the CC strains. Distance between bacterial communities measured by Bray–Curtis and unweighted UniFrac metrics was used to analyze statistical significance between CC strains. Analysis of similarity (ANOSIM) and permutational multivariate analysis of variance (PERMANOVA) was performed with 999 permutations and the R and Pseudo-F test statistics are reported along with the P-values

We also noted increased spread within some of the CC strains in the PCOA plot. To determine the effect of individual variance within a CC strain, distances within a strain and across different strains were calculated and plotted (Fig. 2C, D). For both the metrics, within strain distances were significantly smaller compared with distances across the strains (p < 0.05 for Wilcoxon signed-rank test). This analysis showed that the individual variance within a strain had less impact on the significant differences observed in the microbial composition across different CC strains.

To determine the effect of littermates, we performed an analysis using Adonis. Initial litter cages had a small effect (6%) on microbiome composition, but this effect was negligible compared to the effect attributable to the CC strain itself (65%).

Microbiome composition is a heritable trait

Broad-sense heritability was calculated for microbiome composition at the genus level. We assumed that the observed phenotype, microbial composition, is a result of unobserved genetic and environmental factors. The within-strain variance was used to calculate the environmental component and the variance across the CC strains used provided the total variance. The difference between the total and environmental variance produced the genetic component. Heritability was obtained by dividing the genetic variance by the total variance. For the genus-level data, the average heritability across the CC strain was 0.30 (Table S2). Rikenellaceae, Acetatifactor, and Prevotellaceae were the most heritable genera with scores more than 0.60 while Turicibacter and Escherichia-Shigella were the least heritable group. When sex was included as a variable, the average heritability increased to 0.39. Almost all the genera had a positive increase in scores showing that with a strain animals grouped better by sex. The strong heritability scores suggest the presence of genetic control in gut microbiome composition.

QTL analysis reveals 17 significant associations

Differential abundance analysis was performed by analysis of composition of microbiomes (ANCOM). This analysis considers the relative abundances and structural zeros to greatly reduce false discovery rates. At the genus level, ANCOM identified 21 genera that were differentially abundant between the CC strains (Table S3). A heat map representing the relative abundance of genera identified by ANCOM analysis shows the diversity across CC strains (Fig. S2).

Genus-level abundance was considered a physical trait and QTL analysis was performed using R/qtl2. After grouping bacterial abundance by CC strain, any genera that were not present in at least four different strains were removed. Sex and kinship between the animals were considered covariates for the analysis. Peaks that crossed the 85% significant LOD score for individual genera were considered significant (Table S4). QTL analysis resulted in 17 statistically significant QTL peaks on 9 different chromosomes (Table 3). QTL peaks were named “Microbial Abundance in Feces” (Micabf) and numbered from 1 to 17. The heritability scores for these genera were all positive with an average of 0.31. The 1.8 peak drop confidence interval was calculated for each peak. The average length of QTL regions was 6.77 Mb, consistent with previous studies involving CC population [34, 62]. Genus Peptococcus and Alloprevotella had QTLs on two and three different chromosomes respectively. Genera Muribaculaceae (Micabf3) and Erysipelotrichaceae (Micabf9) were associated with regions adjacent to each other on chromosome 2. Similarly, genus Ruminococcaceae (Micabf14) and genus Peptococcus (Micabf16) were associated with regions adjacent to each other on chromosome 9.

Table 3 Analysis of bacterial abundance identifies numerous QTL. Bacterial abundance in the feces of 28 CC strains was used to identify QTL using rQTL2. 17 statistically significant peaks were identified on 9 different chromosomes. Chr, chromosome; Pos, position; LOD, logarithm of the odds; H2, broad sense heritability; proximal and distal regions are in Mbp

Genes within the QTL regions are associated with important functions

We wanted to utilize the millions of SNPs present in the CC genome to further shortlist and correlate the genes to bacterial abundance in the microbiome of different mouse strains. The Mouse Genome Informatics (MGI) database was used to collect all the protein-coding genes and other features from each QTL interval (Table S5). We picked genus Lachnospiraceae UCG-006, which was differentially abundant between the healthy and sick groups after infection, for further study. The QTL plot for this genus displayed a significant peak on chromosome 12 (Fig. 3A). This 6 Mb region contains 45 protein-coding genes. The founder effect plot (Fig. 3B) suggested that CC strains with NOD and 129S1 allele in this region had lower abundance of genus Lachnospiraceae UCG-006 compared to other founder strains. These two alleles could act independently, or the same variant could be causing this effect in different CC strains.

Fig. 3
figure 3

Genus Lachnospiraceae has a significant peak on chromosome 12. A LOD plot for rank transformed relative abundance values for genus Lachnospiraceae. The dotted green line represents 85% significance and blue line represents 90% significance. B Founder effect plot for genus Lachnospiraceae on chromosome 12

We shortlisted all variants that matched this founder allele pattern and identified 1490 variants representing 31 of the 45 protein-coding genes. Then, we used the variant effect predictor to identify variants that had high impact on the protein structure. This analysis identified three genes with missense mutations and one gene with a 3′ UTR variant (Table 4). Which of these high-impact mutations is involved in determining the composition of genus Lachnospiraceae in the CC strains requires further experimentation. We repeated this process for all the other significant peaks and found high-impact variants for 11 regions (Table S6).

Table 4 Top SNP’s for genus Lachnospiraceae on chromosome 12. Founder allele pattern was used to filter snp’s within the significant interval in chromosome 12. Variant effect predictor was used to calculate the consequence for the variants

In order to identify pathways associated with the variants we identified, we used the Genecards database to identify putative functions for high-impact variants. Several of these genes, including Ralbp1 [63], Pfas [64], Tyro3 [65], Opcml [66, 67], Syne2 [68], Six4 [69], Tll2 [70], and Slc23a2 [65], were previously implicated in gut microbiome abundance, metabolic traits, and several intestinal diseases. A comprehensive list of all the gene summaries, previous GWAS analysis, and pathways can be found in Table S7.

We further analyzed these genes for enrichment using enrichr. This database has more than 400,000 terms from 200 different libraries. Any term with a p value of < 0.05 was considered significant. These genes were enriched for terms associated with BMI, metabolic traits, immune response, biochemical pathways, and neurological conditions (Table S8). Association of these variants with significant biological processes suggests the importance of these genes in determining the composition of the gut microbiome.

Differential abundance of genera Parasutterella and Lachnospiraceae-UCG-006 correlated with STm infection outcome

A subset of our CC strains were infected with Salmonella enterica serotype Typhimurium (STm) 2 days after the collection of feces for this study. We wanted to determine if the abundance of different genera influenced the infection outcome. Forty-eight animals from 8 different CC strains, 3 males and 3 females per strain, were included in this analysis. Four strains with a median survival time of less than 7 days were classified as “sick” and the other four strains that survived to day 7 post-STm infection without clinical signs were classified as “healthy.” ANCOM analysis on pre-infection feces between these two groups was performed after reducing the counts to the genus level. This analysis identified two genera as differentially abundant in the pre-infection microbiomes of healthy and sick mice (Fig. 4A and Table S3). Increased abundance of genus Lachnospiraceae UCG-006 and decreased abundance of genus Parasutterella were correlated with healthy outcome after infection with STm (Fig. 4B, C).

Fig. 4
figure 4

ANCOM identifies genera that are differentially abundant in healthy versus sick strains. Genera (L6) differentially abundant between healthy and sick groups were identified by ANCOM. W-Statistic score above 0.7 was considered significant (A). Genus Lachnospiraceae UCG-006, class Clostridia, phylum Firmicutes, is present in higher abundance in CC strains that remained healthy after STm infection (B). Genus Parasutterella, class Betaproteobacteria, phylum Proteobacteria, were in lower abundance in animals that remained healthy after STm infection (C). P < 0.05 (*) and P < 0.001 (***) for Student’s t-test

Machine learning accurately predicts the metadata groups

We used bacterial composition to predict the metadata groups (CC strain and healthy vs. sick after STm infection) using a machine learning classifier. The bacterial composition data at the genus level was randomly split into training and test sets, and the training set was used to train a random forest classifier. Then, each test sample was used to predict the metadata group based on its fecal microbial composition. Cross-validations were performed during parameter optimization and feature selection to tune the model.

First, for the CC strain data, the classifier assigned blinded test data to the right CC strain group with 90% accuracy and a cross-validation average accuracy of 0.89. The mislabeled CC strain groups were mostly random (Fig S3). This mislabeling may be due to the small sample size used (n = 5) when training the classifier. The top features identified from the model were genera Parabacteroides, Rikenellaceae, and Odoribacter (Table S9). The importance score of the top 5 features from the model added up to only 0.13 again illustrating diversity in the microbiome composition across the CC strains. Second, for the STm infection outcome data, the classifier mislabeled some of the healthy animals into the sick group leading to an overall accuracy of 90% and an AUROC of 0.97 for the model (Fig. 5A, B). Genera Lachnospiraceae UCG-006 and Parasutterella, which were also identified by ANCOM, along with genus Preveotellaceae_UCG-001, were the top features for this classifier. The importance of the top 5 features added up to 0.26. This higher value indicates that infection outcome prediction is more tightly controlled by fewer genera than the strain classifier (Table S9). The higher accuracy rate of these classifiers shows that machine learning can be successfully used in predicting both strain and infection outcome using the microbiome data.

Fig. 5
figure 5

Machine learning algorithm predicts infection outcome. The area under the random operation curve (AUROC) plot for the trained random forest health classifier. AUC (area under the curve) = 0.97 (A). Model accuracy plot for the classifier (B)

Discussion

Growing evidence suggests a complex interaction between host genes and the microbiome [26, 29, 71, 72]. Furthermore, the combined effects of host genetics and microbiome composition on several phenotypes including sleep, anxiety, and liver damage have previously been identified [34, 36, 37]. The effect of diet, environment, early exposure to certain microbes, and the use of antibiotics makes it challenging to study these interactions in humans.

We analyzed the bacterial composition of the feces collected from 28 genetically diverse strains of CC mice. Despite identical housing, food, and husbandry, fecal microbial communities differed significantly across CC strains, beginning at the phylum level. The most abundant phylum, Bacteroidetes, had a relative abundance that varied from 40 to 80% across CC strains. The second most abundant phylum, the Firmicutes, had a relative abundance that varied between 17 and 58% across CC strains (Table 1). But Firmicutes were the most abundant phylum in the previously published CC cecal microbiome data [34, 73]. This difference is likely due to sample collection site differences (feces vs. cecum) [34, 73]. Comparing all the datasets in the Murine Microbiome Database, Firmicutes were the most abundant phylum in the cecum (70%), while Bacteroidetes were the most abundant phylum in the feces (46%) [74]. In our study, the other eight observed phyla were absent in as few as one strain and as many as 26 of the 28 strains. This difference across the CC strains was also noticeable in the alpha diversity metric, Shannon’s index (Fig. 1B). Diversity starting at such a high level is interesting and warrants consideration of the microbiome as an important factor when comparing phenotypic outcomes from different inbred strains.

Animals within a CC strain tended to have more similar bacterial composition profiles in the feces than animals from different CC strains. The PCOA plots for two of the diversity metrics (Bray–Curtis and unweighted UniFrac) grouped mice better by strain than by any other observed parameter. PERMANOVA and ANSOIM analysis further confirmed significant differences across CC strains. Next, we tested whether these differences were caused by the variations in individual animals within a strain. We calculated the distance in individuals within a strain and between strains. For both diversity metrics, the between-strain distances were significantly higher than the individual distance within the strains. Similar observation was made in the microbiome study involving CC founders [73]. This close association of bacteria within a strain and diversity across the strain makes CC mice a great tool for studying the effect of host genetics on the gut microbiome.

Keeping in mind the limitations of 16S rRNA gene sequencing in prediction of species-level data, we identified genetic associations with the abundance of certain microbes at the genus level. All genera had positive heritability scores with more than half of the genera having a score of 0.25 and above. These broad sense heritability values do not take the founder genotype into account. When sex was included as a variable, heritability scores increased suggesting tighter control by sex within CC strains. Given the influence of environment and diet on gut microbiome, these values suggest that microbiome composition has a considerable heritability [75].

We also employed a machine-learning algorithm to predict the effect of host genotype on the microbiome. The Random Forest classifier had an overall accuracy of 90% and cross-validation score of 0.89 in matching microbiome composition to the correct CC strain. The accuracy of the classifier to properly predict the CC strain based on the genus-level bacterial composition data provides further evidence for the influence of host genetics on microbial abundance.

For successful genetic association and QTL mapping, a phenotype must be diverse across the CC strains. Using R/qtl2 analysis, we identified 17 significant peaks on 9 different chromosomes linked to 14 different genera. The 1.8 peak drop confidence interval ranged from 0.3 to 26 Mb with a median range of 4.36 Mb. These short QTL regions, given the size of the population included in this study, demonstrate the power of QTL mapping with CC strains.

Several similar studies in mice have identified QTL linked to the presence of particular members of the microbiota. Micabf6, a region on chromosome 10 that we identify as linked to the presence of genus Alloprevotella, overlapped with six QTL regions identified in the Advanced Intercross population as linked to particular members of the microbiota. These QTL regions were associated with the presence of Coriobacteriales, Streptococcaceae, and Lactococcus [29]. Micaf8 on chromosome 9 also overlaps with a QTL identified from the Advanced intercross population for genus Barnesiella. Micabf11 on chromosome 17 associated with the presence of Lachnospiraceae in our study, matches three different QTL regions identified in the BXD population study [31]. These three regions were associated with the presence of Bacillales and Staphylococcus.

The only other QTL study involving CC strains [34], examined microbial abundance in the cecum. Three QTL associated with microbial matched QTL identified in this study. Genetic regions associated with genera Caminicella, Turicibacter, and Tannerella overlapped with Micabf8 (Chr 9, Erysipelotrichaceae), Micabf7 (Chr 10, Mucispirillum), and Micabf1 (Chr 19, Enterorhabdus), respectively. Even though the genera associated with each of these overlapping regions did not match, these regions may be influencing microbial abundance by a common mechanism.

For individual QTL identified in this study, we looked closely into the genes with structural variants in each interval. Many genes in the QTL we identified had critical roles in important body functions including metabolism and immunity (Table S7). KEG pathway enrichment analysis for these genes picked several important metabolic pathways including protein absorption, sucrose metabolism, and inositol phosphate metabolism. Similarly, mammalian phenotype enrichment identified genes that are involved in colon morphology, as well as several immune and enzyme modulators. We further shortlisted top candidate genes based on a previous association with human disease (Table 5). Most of these genes were previously associated with gut microbiota composition. These intervals also contained genes associated with body mass index (BMI), intestinal bowel disease (IBD), and ulcerative colitis (UC), suggesting a function for these genes in the intestinal tract [58].

Table 5 Top Genes of Interest for all QTL regions. Shortlisted protein-coding genes based on founder allele pattern, consequence, and previous GWAS hits

Given their location in the intestine and their ability to produce and modify several metabolites, various bacterial species have been proposed or used as probiotics for conditions including obesity, diarrhea, diabetes, Clostridium difficile infection, IBD, and neurological diseases [76,77,78,79,80,81]. Many of these probiotics have failed to produce the desired results at the population level [82,83,84,85,86]. The complexity of and interplay between host genetics and microbial composition may explain why simple supplementation with a given bacterial species has not been particularly successful in genetically diverse populations such as humans.

A decrease in beneficial microbes, enrichment of pathogens, and imbalances of metabolites produced by microbes can alter the outcome of diseases [87]. We wanted to identify correlations between pre-infection bacterial composition and the outcome of STm infection. Previous murine studies have successfully employed machine-learning algorithms to predict the outcome of phenotypes including memory, anxiety, and AOM-induced toxicity [35,36,37]. We employed a similar random forest classifier to predict the CC genotype with respect to microbial composition. The algorithm we used also identified several genera that were important in predicting the outcome of infection with STm. Genus Parasutterella was the top feature that allowed the algorithm to differentiate between animals that remained healthy versus those that became ill after STm infection. To further shortlist genera that were differentially abundant between the sick and healthy groups, we employed ANCOM. This analysis identified Lachnospiraceae UCG-006 in addition to Parasutterella as significantly different between the two groups.

Parasutterella is also found in other host species including humans, rats, dogs, pigs, chicken, turkeys, and calves [88]. Changes in the relative abundance of this genus have been reported in several diseases. Parasutterella are increased in submucosal tissues of patients with advanced Crohn’s disease [89] and are associated with pancreatitis in rats [90]. Increased abundance of Parasutterella is also associated with depression and major depressive disorder [91, 92]. Furthermore, increased abundance of Parasutterella was linked to the genesis and development of irritable bowel syndrome (IBS) and is associated with chronic intestinal inflammation in patients with IBS [93].

Parasutterella produce succinate as a fermentative end product and also alter the production of several microbial-derived metabolites involved in bile acid maintenance, tyrosine, tryptophan, and cholesterol metabolism [88]. In our experiments, genus Parasutterella was abundant in the pre-infection fecal microbiota of CC strains that developed clinical signs after infection with STm. Opportunistic pathogens including, Enterohemorrhagic Escherichia coli and Clostridium difficile, sense succinate produced by the commensal gut microbiota through a transcriptional regulator, catabolite repressor/activator (cra) [94, 95]. STm also utilizes succinate produced by intestinal commensals [96] and cra gene activation is important for STm pathogenesis [97]. Sensing of succinate results in the activation of STm virulence genes, including activating the Salmonella pathogenicity island 2 type III secretion system (SPI2 T3SS), leading to increased pathogenicity [98, 99]. Thus, succinate produced by Parasutterella may be sensed by STm leading to increased virulence and development of severe clinical symptoms in CC strains that harbor this organism in the intestinal tract. This hypothesis remains to be tested.

The family Lachnospiraceae is a core member of the gut microbiota in both mice and humans [74, 100]. Despite being one of the main producers of short-chain fatty acids (SCFA) in the intestine and helping with metabolism [101], the role of family Lachnospiraceae is controversial. The relative abundance of multiple genera in this family can both positively and negatively influence several diseases, including obesity, diabetes, IBD, and depressive syndrome [102]. In our study, the relative abundance of genus Lachnospiraceae UCG-006 is associated with a positive health outcome after STm infection. Increased relative abundance of genus Lachnospiraceae UCG-006 is correlated with positive outcomes in several diseases including colon cancer, IBD, LPS-induced inflammation, and colitis ([103,104,105,106]. Butyrate, a SCFA, can be utilized by intestinal epithelial cells as an energy source [107, 108]. Butyrate also reduces the colonization and virulence of several Salmonella sp. including Typhimurium [109,110,111,112,113]. It is possible that SCFA produced by Lachnospiraceae species may reduce the initial intestinal colonization of STm, leading to a positive outcome after infection.

Conclusions

Diversity and homeostasis are key to the proper functioning of our gut microbiome. Changes in this delicate balance can lead to metabolic abnormalities, intestinal inflammation, infection by pathogens, autoimmunity, and neurological diseases. By controlling the diet, environment, and varying host genetics, we identified multiple regions of the mouse genome that are associated with intestinal colonization by specific microbes. Despite the small sample size, we successfully used machine learning tools to predict the metadata columns. With publicly available bacterial composition data and advancements in the artificial intelligence field, more machine learning tools can be employed to develop microbes as biomarkers for disease prediction. Future genetic modification and gnotobiotic studies in murine models such as the CC panel will be useful to convert the correlated genes to causal candidates for therapeutic application. A clear understanding of the role of environment, diet, and host genetics will provide a basis for using microbes as a personalized therapeutic tool to prevent, diagnose and treat various body conditions and diseases.

Availability of data and materials

All the raw data files and supplemental information is publicly available at https://figshare.com/articles/dataset/Raw_FastQ_files/21964754 and https://figshare.com/articles/journal_contribution/Supplemental_Information/21964745. ASVs obtained for individual mice along with the metadata are available in Table S10. QTL, Founder effect, and Genotype vs phenotype plots for all the significant QTLs are available as supplementary file (Fig S4).

References

  1. Ley RE, Peterson DA, Gordon JI. Ecological and evolutionary forces shaping microbial diversity in the human intestine. Cell. 2006;124(4):837–48.

    Article  CAS  PubMed  Google Scholar 

  2. Whitman WB, Coleman DC, Wiebe WJ. Prokaryotes: the unseen majority. Proc Natl Acad Sci. 1998;95(12):6578–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Shreiner AB, Kao JY, Young VB. The gut microbiome in health and in disease. Curr Opin Gastroenterol. 2015;31(1):69–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Perry RJ, et al. Acetate mediates a microbiome–brain–β-cell axis to promote metabolic syndrome. Nature. 2016;534(7606):213–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Bunyavanich S, Berin MC. Food allergy and the microbiome: Current understandings and future directions. J Allergy Clin Immunol. 2019;144(6):1468–77.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Cuív PÓ, et al. The gut bacterium and pathobiont Bacteroides vulgatus activates NF-κB in a human gut epithelial cell line in a strain and growth phase dependent manner. Anaerobe. 2017;47:209–17.

    Article  Google Scholar 

  7. Vuong HE, Hsiao EY. Emerging roles for the gut microbiome in autism spectrum disorder. Biol Psychiatry. 2017;81(5):411–23.

    Article  PubMed  Google Scholar 

  8. Lubomski M, et al. Parkinson’s disease and the gastrointestinal microbiome. J Neurol. 2020;267(9):2507–23.

    Article  PubMed  Google Scholar 

  9. Lin L, Zheng LJ, Zhang LJ. Neuroinflammation, gut microbiome, and Alzheimer’s disease. Mol Neurobiol. 2018;55(11):8243–50.

    Article  CAS  PubMed  Google Scholar 

  10. Dahlin M, Prast-Nielsen S. The gut microbiome and epilepsy. EBioMedicine. 2019;44:741–6.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Benakis C, et al. The microbiome-gut-brain axis in acute and chronic brain diseases. Curr Opin Neurobiol. 2020;61:1–9.

    Article  CAS  PubMed  Google Scholar 

  12. Li Y, et al. The role of microbiome in insomnia, circadian disturbance and depression. Front Psychiatry. 2018;9:669.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Churchill GA, et al. The collaborative cross, a community resource for the genetic analysis of complex traits. Nat Genet. 2004;36(11):1133.

    Article  CAS  PubMed  Google Scholar 

  14. Threadgill DW, Churchill GA. Ten years of the collaborative cross. Genetics. 2012;190(2):291–4.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Morgan AP, Welsh CE. Informatics resources for the collaborative cross and related mouse populations. Mamm Genome. 2015;26(9–10):521–39.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Graham JB, et al. Genetic diversity in the collaborative cross model recapitulates human West Nile virus disease outcomes. MBio. 2015;6(3).

  17. Rasmussen AL, et al. Host genetic diversity enables Ebola hemorrhagic fever pathogenesis and resistance. Science. 2014;346(6212):987–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Gu BAO, et al. Collaborative Cross mice reveal extreme epilepsy phenotypes and genetic loci for seizure susceptibility. LID. 2020(1528–1167 (Electronic)). https://doi.org/10.1111/epi.16617

  19. Abu-Toamih Atamni HJ, et al. Glucose tolerance female-specific QTL mapped in collaborative cross mice. Mamm Genome. 2017;28(1–2):20–30.

    Article  CAS  PubMed  Google Scholar 

  20. Smith CM, et al. Functionally overlapping variants control tuberculosis susceptibility in Collaborative Cross mice. mBio. 2019;10(6).

  21. Noll KE, et al. Complex genetic architecture underlies regulation of influenza-A-virus-specific antibody responses in the Collaborative Cross. Cell Rep. 2020;31(4):107587.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Lewis L, et al. Population-based analysis of DNA damage and epigenetic effects of 1,3-butadiene in the mouse. Chem Res Toxicol. 2019;32(5):887–98.

    Article  CAS  PubMed  Google Scholar 

  23. Scoggin K, et al. Genetic background influences survival of infections with <em>Salmonella enterica</em> serovar Typhimurium in the Collaborative Cross. Accepted for publication in PLoS Genetics. 2022:2022.02.07.479341.

  24. Zhang J, et al. Identification of new loci involved in the host susceptibility to Salmonella Typhimurium in collaborative cross mice. BMC Genomics. 2018;19(1):1–13.

    Article  Google Scholar 

  25. Dąbrowska K, Witkiewicz W. Correlations of host genetics and gut microbiome composition. Front Microbiol. 2016;7:1357.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Goodrich JK, et al. Human genetics shape the gut microbiome. Cell. 2014;159(4):789–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Turnbaugh PJ, et al. A core gut microbiome in obese and lean twins. Nature. 2009;457(7228):480–4.

    Article  CAS  PubMed  Google Scholar 

  28. Yatsunenko T, et al. Human gut microbiome viewed across age and geography. Nature. 2012;486(7402):222–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Benson AK, et al. Individuality in gut microbiota composition is a complex polygenic trait shaped by multiple environmental and host genetic factors. Proc Natl Acad Sci U S A. 2010;107(44):18933–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Leamy LJ, et al. Host genetics and diet, but not immunoglobulin A expression, converge to shape compositional features of the gut microbiome in an advanced intercross population of mice. Genome Biol. 2014;15(12):552.

    Article  PubMed  PubMed Central  Google Scholar 

  31. McKnite AM, et al. Murine gut microbiota is defined by host genetics and modulates variation of metabolic traits. PLoS ONE. 2012;7(6):e39191.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Org E, et al. Genetic and environmental control of host-gut microbiota interactions. Genome Res. 2015;25(10):1558–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Snijders AM, et al. Influence of early life exposure, host genetics and diet on the mouse gut microbiome and metabolome. Nat Microbiol. 2016;2(2):1–8.

    Article  Google Scholar 

  34. Bubier JA, et al. A microbe associated with sleep revealed by a novel systems genetic analysis of the microbiome in Collaborative Cross mice. Genetics. 2020;214(3):719–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Mao JH, et al. Genetic and metabolic links between the murine microbiome and memory. Microbiome. 2020;8(1):53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Jin X, et al. Gut microbiome partially mediates and coordinates the effects of genetics on anxiety-like behavior in Collaborative Cross mice. Sci Rep. 2021;11(1):270.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Zhong C, et al. Host genetics and gut microbiota cooperatively contribute to azoxymethane-induced acute toxicity in Collaborative Cross mice. Arch Toxicol. 2021;95(3):949–58.

    Article  CAS  PubMed  Google Scholar 

  38. Bogomolnaya LM, et al. ‘Form variation’of the O12 antigen is critical for persistence of Salmonella Typhimurium in the murine intestine. Mol Microbiol. 2008;70(5):1105–19.

    Article  CAS  PubMed  Google Scholar 

  39. Bolyen E, et al. Reproducible, interactive, scalable and extensible microbiome data science using QIIME 2. Nat Biotechnol. 2019;37(8):852–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Callahan BJ, et al. DADA2: High-resolution sample inference from Illumina amplicon data. Nat Methods. 2016;13(7):581–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Bokulich NA, et al. Optimizing taxonomic classification of marker-gene amplicon sequences with QIIME 2’s q2-feature-classifier plugin. Microbiome. 2018;6(1):90.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Quast C, et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res. 2013;41(Database issue):D590–6.

    CAS  PubMed  Google Scholar 

  43. Katoh K, et al. MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 2002;30(14):3059–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Price MN, Dehal PS, Arkin AP. FastTree 2–approximately maximum-likelihood trees for large alignments. PLoS ONE. 2010;5(3):e9490.

    Article  PubMed  PubMed Central  Google Scholar 

  45. Lozupone C, Knight R. UniFrac: a new phylogenetic method for comparing microbial communities. Appl Environ Microbiol. 2005;71(12):8228–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Bray JR, Curtis JT. An Ordination of the Upland Forest Communities of Southern Wisconsin. Ecol Monogr. 1957;27(4):325–49.

    Article  Google Scholar 

  47. Anderson MJ. A new method for non-parametric multivariate analysis of variance. Austral Ecol. 2001;26(1):32–46.

    Google Scholar 

  48. Clarke K. Nonparametric multivariate analyses of changes in community structure. Austral Ecol. 1993;18:117–43.

    Article  Google Scholar 

  49. Mandal S, et al. Analysis of composition of microbiomes: a novel method for studying microbial composition. Microb Ecol Health Dis. 2015;26:27663.

    PubMed  Google Scholar 

  50. Lin H, Peddada SD. Analysis of compositions of microbiomes with bias correction. Nat Commun. 2020;11(1):3514.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Collin R, et al. Common heritable immunological variations revealed in genetically diverse inbred mouse strains of the Collaborative Cross. J Immunol. 2019;202(3):777–86.

    Article  CAS  PubMed  Google Scholar 

  52. Visscher PM, Hill WG, Wray NR. Heritability in the genomics era–concepts and misconceptions. Nat Rev Genet. 2008;9(4):255–66.

    Article  CAS  PubMed  Google Scholar 

  53. Broman KW, et al. R/qtl2: Software for mapping quantitative trait loci with high-dimensional data and multiparent populations. Genetics. 2019;211(2):495–502.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Vincent M, et al. QTLViewer: an interactive webtool for genetic analysis in the Collaborative Cross and Diversity Outbred mouse populations. G3 Genes Genom Genet. 2022;12(8).

  55. Blake JA, et al. Mouse Genome Database (MGD): Knowledgebase for mouse-human comparative biology. Nucleic Acids Res. 2021;49(D1):D981-d987.

    Article  CAS  PubMed  Google Scholar 

  56. Broman KW. SQLite database of variants in Collaborative Cross founder mouse strains. 2017.

    Google Scholar 

  57. McLaren W, et al. The Ensembl variant effect predictor. Genome Biol. 2016;17(1):122.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Stelzer G, et al. The GeneCards suite: from gene data mining to disease genome sequence analyses. Curr Prot Bioinform. 2016;54(1):1.30.1-1.30.33.

    Google Scholar 

  59. Chen EY, et al. Enrichr: interactive and collaborative HTML5 gene list enrichment analysis tool. BMC Bioinformatics. 2013;14:128.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Kuleshov MV, et al. Enrichr: a comprehensive gene set enrichment analysis web server 2016 update. Nucleic Acids Res. 2016;44(W1):W90–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Pedregosa F, et al. Scikit-learn: machine learning in Python. J Mach Learn Res. 2011;12(1):2825–30.

    Google Scholar 

  62. Snijders AM, et al. Influence of early life exposure, host genetics and diet on the mouse gut microbiome and metabolome. Nat Microbiol. 2016;2:16221.

    Article  CAS  PubMed  Google Scholar 

  63. Scepanovic P, et al. A comprehensive assessment of demographic, environmental, and host genetic associations with gut microbiome diversity in healthy individuals. Microbiome. 2019;7(1):130.

    Article  PubMed  PubMed Central  Google Scholar 

  64. Christakoudi S, et al. GWAS of allometric body-shape indices in UK Biobank identifies loci suggesting associations with morphogenesis, organogenesis, adrenal cell renewal and cancer. Sci Rep. 2021;11(1):10688.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Rühlemann MC, et al. Genome-wide association study in 8,956 German individuals identifies influence of ABO histo-blood groups on gut microbiome. Nat Genet. 2021;53(2):147–55.

    Article  PubMed  Google Scholar 

  66. Hughes DA, et al. Genome-wide associations of human gut microbiome variation and implications for causal inference analyses. Nat Microbiol. 2020;5(9):1079–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Ishida S, et al. Genome-wide association studies and heritability analysis reveal the involvement of host genetics in the Japanese gut microbiota. Commun Biol. 2020;3(1):686.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Lopera-Maya EA, et al. Effect of host genetics on the gut microbiome in 7,738 participants of the Dutch Microbiome Project. Nat Genet. 2022;54(2):143–51.

    Article  CAS  PubMed  Google Scholar 

  69. Fan W, et al. Association between human genetic variants and the vaginal bacteriome of pregnant women. mSystems. 2021;6(4):e0015821.

    Article  PubMed  Google Scholar 

  70. Qin Y, et al. Combined effects of host genetics and diet on human gut microbiota and incident disease in a single population cohort. Nat Genet. 2022;54(2):134–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Blekhman R, et al. Host genetic variation impacts microbiome composition across human body sites. Genome Biol. 2015;16(1):191.

    Article  PubMed  PubMed Central  Google Scholar 

  72. Chen C, et al. Contribution of host genetics to the variation of microbial composition of cecum lumen and feces in pigs. Front Microbiol. 2018;9:2626.

    Article  PubMed  PubMed Central  Google Scholar 

  73. Campbell JH, et al. Host genetic and environmental effects on mouse intestinal microbiota. ISME J. 2012;6(11):2033–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Yang J, et al. Introducing Murine Microbiome Database (MMDB): a curated database with taxonomic profiling of the healthy mouse gastrointestinal microbiome. Microorganisms. 2019;7(11).

  75. Odet F, et al. The founder strains of the Collaborative Cross express a complex combination of advantageous and deleterious traits for male reproduction. G3 (Bethesda, Md). 2015;5(12):2671–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Hempel S, et al. Probiotics for the prevention and treatment of antibiotic-associated diarrhea: a systematic review and meta-analysis. JAMA. 2012;307(18):1959–69.

    Article  CAS  PubMed  Google Scholar 

  77. Goldenberg JZ, et al. Probiotics for the prevention of Clostridium difficile-associated diarrhea in adults and children. Cochrane Database Syst Rev. 2017;12(12):CD006095. https://doi.org/10.1002/14651858.CD006095.pub4.

  78. Ford AC, et al. Systematic review with meta-analysis: the efficacy of prebiotics, probiotics, synbiotics and antibiotics in irritable bowel syndrome. Aliment Pharmacol Ther. 2018;48(10):1044–60.

    Article  PubMed  Google Scholar 

  79. Luoto R, et al. Impact of maternal probiotic-supplemented dietary counselling on pregnancy outcome and prenatal and postnatal growth: a double-blind, placebo-controlled study. Br J Nutr. 2010;103(12):1792–9.

    Article  CAS  PubMed  Google Scholar 

  80. Westfall S, et al. Microbiome, probiotics and neurodegenerative diseases: deciphering the gut brain axis. Cell Mol Life Sci. 2017;74(20):3769–87.

    Article  CAS  PubMed  Google Scholar 

  81. Quigley EMM. Microbiota-Brain-Gut Axis and Neurodegenerative Diseases. Curr Neurol Neurosci Rep. 2017;17(12):94.

    Article  PubMed  Google Scholar 

  82. Allen SJ, et al. Lactobacilli and bifidobacteria in the prevention of antibiotic-associated diarrhoea and Clostridium difficile diarrhoea in older inpatients (PLACIDE): a randomised, double-blind, placebo-controlled, multicentre trial. Lancet. 2013;382(9900):1249–57.

    Article  PubMed  Google Scholar 

  83. Freedman SB, et al. Multicenter trial of a combination probiotic for children with gastroenteritis. N Engl J Med. 2018;379(21):2015–26.

    Article  CAS  PubMed  Google Scholar 

  84. Georgieva M, et al. Use of the probiotic lactobacillus reuteri dsm 17938 in the prevention of antibiotic-associated infections in hospitalized Bulgarian children: a randomized, controlled trial. J IMAB Ann Proc Sci Papers. 2015;21(4):895–900.

    Google Scholar 

  85. Dermyshi E, et al. The “golden age” of probiotics: a systematic review and meta-analysis of randomized and observational studies in preterm infants. Neonatology. 2017;112(1):9–23.

    Article  CAS  PubMed  Google Scholar 

  86. Zhang G-Q, et al. Probiotics for preventing late-onset sepsis in preterm neonates: a PRISMA-compliant systematic review and meta-analysis of randomized controlled trials. Medicine. 2016;95(8).

  87. Duvallet C, et al. Meta-analysis of gut microbiome studies identifies disease-specific and shared responses. Nat Commun. 2017;8(1):1784.

    Article  PubMed  PubMed Central  Google Scholar 

  88. Ju T, et al. Defining the role of Parasutterella, a previously uncharacterized member of the core gut microbiota. Isme j. 2019;13(6):1520–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Chiodini RJ, et al. Microbial population differentials between mucosal and submucosal intestinal tissues in advanced Crohn’s disease of the ileum. PLoS ONE. 2015;10(7):e0134382.

    Article  PubMed  PubMed Central  Google Scholar 

  90. Huang C, et al. Dysbiosis of intestinal microbiota and decreased antimicrobial peptide level in paneth cells during hypertriglyceridemia-related acute necrotizing pancreatitis in rats. Front Microbiol. 2017;8:776.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Barandouzi ZA, et al. Altered composition of gut microbiota in depression: a systematic review. Front Psychiatry. 2020;11:541.

    Article  PubMed  PubMed Central  Google Scholar 

  92. Cheung SG, et al. Systematic review of gut microbiota and major depression. Front Psychiatry. 2019;10:34.

    Article  PubMed  PubMed Central  Google Scholar 

  93. Chen YJ, et al. Parasutterella, in association with irritable bowel syndrome and intestinal chronic inflammation. J Gastroenterol Hepatol. 2018;33(11):1844–52.

    Article  CAS  PubMed  Google Scholar 

  94. Curtis MM, et al. The gut commensal Bacteroides thetaiotaomicron exacerbates enteric infection through modification of the metabolic landscape. Cell Host Microbe. 2014;16(6):759–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Ferreyra JA, et al. Gut microbiota-produced succinate promotes C. difficile infection after antibiotic treatment or motility disturbance. Cell Host Microbe. 2014;16(6):770–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Spiga L, et al. An oxidative central metabolism enables Salmonella to utilize microbiota-derived succinate. Cell Host Microbe. 2017;22(3):291-301.e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Allen JH, et al. A functional cra gene is required for Salmonella enterica serovar typhimurium virulence in BALB/c mice. Infect Immun. 2000;68(6):3772–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Connors J, Dawe N, Van Limbergen J. The role of succinate in the regulation of intestinal inflammation. Nutrients. 2019;11(1):25.

    Article  CAS  Google Scholar 

  99. Rosenberg G, et al. Host succinate is an activation signal for Salmonella virulence during intracellular infection. Science. 2021;371(6527):400–5.

    Article  CAS  PubMed  Google Scholar 

  100. Dai D, et al. GMrepo v2: a curated human gut microbiome database with special focus on disease markers and cross-dataset comparison. Nucleic Acids Res. 2021;50(D1):D777–84.

    Article  PubMed Central  Google Scholar 

  101. Flint HJ, et al. Links between diet, gut microbiota composition and gut metabolism. Proc Nutr Soc. 2015;74(1):13–22.

    Article  CAS  PubMed  Google Scholar 

  102. Vacca M, et al. The controversial role of human gut Lachnospiraceae. Microorganisms. 2020;8(4):573.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Guo M, Li Z. Polysaccharides isolated from Nostoc commune Vaucher inhibit colitis-associated colon tumorigenesis in mice and modulate gut microbiota. Food Funct. 2019;10(10):6873–81.

    Article  CAS  PubMed  Google Scholar 

  104. Guo W, et al. Protective effects of microbiome-derived inosine on lipopolysaccharide-induced acute liver damage and inflammation in mice via mediating the TLR4/NF-κB pathway. J Agric Food Chem. 2021;69(27):7619–28.

    Article  CAS  PubMed  Google Scholar 

  105. Wang YN, et al. Effects of probiotics and prebiotics on intestinal microbiota in mice with acute colitis based on 16S rRNA gene sequencing. Chin Med J (Engl). 2019;132(15):1833–42.

    Article  CAS  PubMed  Google Scholar 

  106. Hu L, et al. Nitrate ameliorates dextran sodium sulfate-induced colitis by regulating the homeostasis of the intestinal microbiota. Free Radic Biol Med. 2020;152:609–21.

    Article  CAS  PubMed  Google Scholar 

  107. Clausen MR, Mortensen PB. Kinetic studies on colonocyte metabolism of short chain fatty acids and glucose in ulcerative colitis. Gut. 1995;37(5):684–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Józefiak D, Rutkowski A, Martin SA. Carbohydrate fermentation in the avian ceca: a review. Anim Feed Sci Technol. 2004;113(1):1–15.

    Article  Google Scholar 

  109. Namkung H, et al. Antimicrobial activity of butyrate glycerides toward Salmonella Typhimurium and Clostridium perfringens. Poult Sci. 2011;90(10):2217–22.

    Article  CAS  PubMed  Google Scholar 

  110. Sunkara LT, et al. Butyrate enhances disease resistance of chickens by inducing antimicrobial host defense peptide gene expression. PLoS ONE. 2011;6(11):e27225.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Schulthess J, et al. The Short Chain Fatty Acid Butyrate Imprints an Antimicrobial Program in Macrophages. Immunity. 2019;50(2):432-445.e7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Chu B, et al. Butyrate-mediated autophagy inhibition limits cytosolic Salmonella Infantis replication in the colon of pigs treated with a mixture of Lactobacillus and Bacillus. Vet Res. 2020;51(1):99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Gupta A, et al. Sodium butyrate reduces Salmonella Enteritidis infection of chicken enterocytes and expression of inflammatory host genes in vitro. Front Microbiol. 2020;11.

Download references

Acknowledgements

We thank the Texas A&M Institute for Genome Sciences and Society (TIGSS) for providing genomic services and Dr. John C Blazier (Assistant Research Scientist – TIGSS) for his help with sequence analysis. We would also like to thank QIMME2 creators and forum administrators for their help with the 16S rRNA gene data analysis and troubleshooting. We also thank Dr. Vince Young, Dr. Christine Bassis, and Dr. Lydia Bogomolnaya for helpful discussions.

Funding

This work was funded by the Defense Advanced Research Project Agency (DARPA), project DARPA D17AP00004, and by internal funding from the Texas A&M University College of Medicine awarded to HLAP and D.W.T.

Author information

Authors and Affiliations

Authors

Contributions

D.W.T. and H.L.A.P. designed the experiments in this work. A.N., K.S., and J.G. performed the experiments. A.N. analyzed the data. A.N., H.A.P., and D.W.T. wrote the manuscript and all authors participated in editing the text. Funding for this project was secured by D.W.T. and H.A.P. The authors read and approved the final manuscript.

Corresponding author

Correspondence to Helene L. Andrews-Polymenis.

Ethics declarations

Ethics approval and consent to participate

Mouse studies were conducted on Animal Care and Use protocol number 2015-0315D approved by the animal care and use committee, and in accordance with the Guide for the Care and Use of Laboratory Animals of the National Institutes of Health.

Consent for publication

The authors consent to publish this work.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1:

Figure S1. Sequence counts significantly decreased after DADA2 processing. (A) Raw sequence counts from Illumina paired end sequencing. (B) Sequence counts after filtering, denoising, chimeric removal and merging using DADA2 pipeline.

Additional file 2:

Figure S2. Diverse genus communities are observed across the CC strains. The heatmap of log transformed relative abundances of top differentially abundant genera between CC strains as identified by ANCOM.

Additional file 3:

Figure S3. Machine learning algorithm accurately predicts the metadata columns. Random forest classifier was trained using the bacterial composition data and predict the CC strain type.

Additional file 4:

Figure S4. Plots for all the significant QTLs. A. LOD plot for rank transformed relative abundance values for significant genera. The dotted lines (Green – 85%, Blue – 90%, Red – 95%) represent significant thresholds calculated by 999 permutations. B. Founder effect plot and LOD plot for the significant chromosome. C. Genotype X Phenotype plot for the highest SNP on the significant chromosome.

Additional file 5:

Table S1. Genera counts and Frequency.

Additional file 6:

Table S2. Broad sense heritability scores.

Additional file 7:

Table S3. ANCOM analysis.

Additional file 8:

Table S4. LOD scores.

Additional file 9:

Table S5. All protein coding genes and mouse QTL's within each significant region.

Additional file 10:

Table S6. Top Snps in each QTL.

Additional file 11:

Table S7. Important Genes, GWAS associations and pathways.

Additional file 12:

Table S8. Enrichment Analysis.

Additional file 13:

Table S9. Machine Learning Importance scores.

Additional file 14:

Table S10. ASVs.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nagarajan, A., Scoggin, K., Gupta, J. et al. Using the collaborative cross to identify the role of host genetics in defining the murine gut microbiome. Microbiome 11, 149 (2023). https://doi.org/10.1186/s40168-023-01552-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40168-023-01552-8