Skip to main content
eLife logoLink to eLife
. 2016 Nov 22;5:e19130. doi: 10.7554/eLife.19130

Cell culture-based profiling across mammals reveals DNA repair and metabolism as determinants of species longevity

Siming Ma 1, Akhil Upneja 1, Andrzej Galecki 2,3,4, Yi-Miau Tsai 2,3, Charles F Burant 5, Sasha Raskind 5, Quanwei Zhang 6, Zhengdong D Zhang 6, Andrei Seluanov 7, Vera Gorbunova 7, Clary B Clish 8, Richard A Miller 2,3, Vadim N Gladyshev 1,*
Editor: Amy J Wagers9
PMCID: PMC5148604  PMID: 27874830

Abstract

Mammalian lifespan differs by >100 fold, but the mechanisms associated with such longevity differences are not understood. Here, we conducted a study on primary skin fibroblasts isolated from 16 species of mammals and maintained under identical cell culture conditions. We developed a pipeline for obtaining species-specific ortholog sequences, profiled gene expression by RNA-seq and small molecules by metabolite profiling, and identified genes and metabolites correlating with species longevity. Cells from longer lived species up-regulated genes involved in DNA repair and glucose metabolism, down-regulated proteolysis and protein transport, and showed high levels of amino acids but low levels of lysophosphatidylcholine and lysophosphatidylethanolamine. The amino acid patterns were recapitulated by further analyses of primate and bird fibroblasts. The study suggests that fibroblast profiling captures differences in longevity across mammals at the level of global gene expression and metabolite levels and reveals pathways that define these differences.

DOI: http://dx.doi.org/10.7554/eLife.19130.001

Research Organism: Human, Mouse, Rat, Rhesus macaque, Other

Introduction

The maximum lifespan of mammalian species differs by more than 100-fold, ranging from ~2 years in shrews to >200 years in bowhead whales (Tacutu et al., 2013). While it has long been observed that maximum lifespan tends to correlate positively with body mass and time to maturity, but negatively with growth rate, mass-specific metabolic rate, and number of offspring (Peters, 1986; Sacher, 1959; Western, 1979), the underlying molecular basis is only starting to be understood.

One way to study the control of longevity is to identify the genes, pathways, and interventions capable of extending lifespan or delaying aging phenotypes in experimental animals. Studies using model organisms have uncovered several important conditions, such as knockout of insulin-like growth factor 1 (IGF-1) receptor (Friedman and Johnson, 1988; Holzenberger et al., 2003; Tatar et al., 2001), inhibition of mechanistic target of rapamycin (mTOR) (Harrison et al., 2009; Kenyon, 2010; Miller et al., 2014), mutation in growth hormone (GH) receptor (Coschigano et al., 2000), ablation of anterior pituitary (e.g. Snell dwarf mice) (Flurkey et al., 2002), augmentation of proteins of the sirtuin family (Chang and Guarente, 2013; Gomes et al., 2013; Mouchiroud et al., 2013; Wood et al., 2004), and restriction of dietary intake (Guarente and Kenyon, 2000; Heilbronn and Ravussin, 2003; McCay et al., 1935; Weindruch et al., 1986). While many of these genes and pathways have been verified in yeast, flies, worms, and mice, the comparisons largely involve treatment and control groups of the same species, and the extent to which they explain the longevity variations across different species is unclear. For example, do the long-lived species have metabolic profiles resembling calorie restriction? Do they suppress IGF-1 or growth hormone signaling compared with the shorter-lived species? More generally, how do the evolutionary strategies of longevity relate to the experimental strategies that extend lifespan in model organisms?

To address these questions, a popular approach has been to compare exceptionally long-lived species with closely related species of common lifespan and identify the features associated with exceptional longevity. Examples include the amino acid changes in Uncoupling Protein 1 (UCP1) and production of high-molecular-mass hyaluronan in the naked mole rat (Kim et al., 2011; Tian et al., 2013); unique sequence changes in IGF1 and GH receptors in Brandt’s bat (Seim et al., 2013); gene gain and loss associated with DNA repair, cell-cycle regulation, and cancer, as well as alteration in insulin signaling in the bowhead whale (Keane et al., 2015; Seim et al., 2014); and duplication of the p53 gene in elephants (Abegglen et al., 2015). Again, it is important to ascertain whether these mechanisms are unique characteristics of specific exceptionally long-lived species, or whether they can also help account for the general lifespan variation (Martin, 1988; Partridge and Gems, 2002).

An extension of this approach has been cross-species analyses in a larger scale. For example, several biochemical studies across multiple mammalian and bird species identified some features correlating with species lifespan. Longevity of fibroblasts and erythrocytes in vitro (Röhme, 1981), poly (ADP-ribose) polymerase activity (Grube and Bürkle, 1992), and rate of DNA repair (Cortopassi and Wang, 1996) were found to be positively correlated with longevity, whereas mitochondrial membrane and liver fatty acid peroxidizability index (Pamplona et al., 1998, 2000), rate of telomere shortening (Haussmann et al., 2003), and oxidative damage to DNA and mitochondrial DNA (Adelman et al., 1988; Barja and Herrero, 2000) showed negative correlation. The advent of high-throughput RNA sequencing (RNAseq) and mass spectrometry technologies has enabled the quantification of whole transcriptomes (Fushan et al., 2015), metabolomes (Ma et al., 2015b), and ionomes (Ma et al., 2015a), across multiple species and organs. These studies revealed the complex transcriptomic and metabolic landscape across different organs and species, as well as some overlaps with the changes observed in the long-lived mutants created in laboratory (Ma et al., 2015b).

While molecular profiling of mammals at the level of tissues may better represent the underlying biology, profiling in cell culture represents more defined experimental conditions and allows further manipulation to alter the identified molecular phenotypes. In this study, we examined the transcriptomes and metabolomes of primary skin fibroblasts across 16 species of mammals, to identify the molecular patterns associated with species longevity. We report that the genes involved in DNA repair and glucose metabolism were up-regulated in the longer lived species, whereas proteolysis and protein translocation activities were suppressed. The longer lived species also had lower levels of lysophosphatidylcholine and lysophosphatidylethanolamine and higher levels of amino acids; and the latter finding was validated in an independent dataset of bird and primate fibroblasts. Thus, molecular insights into longevity may indeed come from defined cell culture systems in mammals.

Results

Gene expression by RNA sequencing

To identify the molecular signatures associating with the differences in longevity, we obtained primary, sun-protected abdominal skin fibroblasts from 13 species of rodents, two species of bats and one species of shrew, representing a wide range of maximum lifespan (ML; from 2.2 years in shrew to 34.0 years in little brown bat) and adult weight (AW; from 10 g in little brown bat to 20 kg in beaver) (Figure 1, Figure 1—source data 1A). Female time to maturity (FTM) and the body mass adjusted residuals (i.e. MLres and FTMres) were included as additional longevity trait (Figure 1, Materials and methods). We profiled gene expression by RNAseq on 28 samples representing 15 species (except gerbil) (Figure 1—source data 1B). Only five of these species had publicly available genomes; this posed a challenge as reliable reference sequences were crucial for accurate RNAseq read alignment and read counting. The gene orthology information was also limited or unavailable for the less common species. To address these issues, we developed a pipeline to obtain species-specific ortholog sets (Figure 2A, Materials and methods). We defined a set of mouse reference sequences based on Ensembl and then performed de novo transcriptome assembly for each species. BLAST was used to find reciprocal best hits between the assembled transcriptome (and published genome, if available) and the mouse reference (Altschul et al., 1997; Camacho et al., 2009; Tatusov et al., 1997). The reciprocal best hits were then trimmed down to open reading frame and the quality of the ortholog sets was assessed by multiple sequence alignment (Materials and methods).

Figure 1. Phylogenetic relationship among species used in the study.

Figure 1.

The tree was constructed using Neighbor-Joining method based on nucleotide sequences. Shrew was used as the out-group. Gerbil was collected for metabolite data only and mouse was included as reference. The species are colored by taxonomic order. Adult Weight (AW), Maximum Lifespan (ML), Female Time to Maturity (FTM), Maximum Lifespan Residual (MLres), and Female Time to Maturity Residual (FTMres) of these species are displayed in log10 scale.

DOI: http://dx.doi.org/10.7554/eLife.19130.002

Figure 1—source data 1. Species and samples used in the current study.
(A) Species and traits information. Life history traits of adult weight (AW, in grams), maximum lifespan (ML, in years), and female time to maturity (FTM, in days) of these species were obtained from Anage database (Tacutu et al., 2013). Since the life history data were not available for meadow vole, the data of a related species Microtus arvalis were used instead. Maximum lifespan residual (MLres) and female time to maturity residuals (FTMres) were computed based on the allometric equations MLres = ML/(4.88×AW0.153) and FTMres = FTM/(78.1×AW0.217), respectively. (B) RNA sequencing and read mapping to ortholog sets. Read mapping statistics are based on STAR. De novo assembly was performed by Trinity. (C) Read mapping to publically available genomes. For the species with publicly available genomes, the reads were also aligned to the full genomes for mapping rate comparison. The numbers of annotated genes were based on the published annotations. (D) Metabolite profiling. Metabolite profiling was performed on selected species only.
DOI: 10.7554/eLife.19130.003

Figure 2. Cross-species analysis of gene expression in cultured skin fibroblasts.

(A) Pipeline to obtain the species-specific ortholog sets and expression values. See Materials and methods or a more detailed description of the methodology. (B) Sequence identity of ortholog sets compared to mouse. Nucleotide and amino acid sequence identity of the ortholog sets in each species was compared to mouse reference (mouse was set as 100%). The ortholog sequences were based on de novo assembled transcriptomes, as well as NCBI genomes (if available; indicated by ‘#’). The box plot shows the distribution across the 9389 gene orthologs, with the central bars indicating median values. (C) Read alignment rates for mapping to complete genomes and ortholog sets. Percent of total reads that could be uniquely aligned to the complete genomes (if available, indicated by ‘#’; shaded bars) or to the ortholog sets are shown. Error bars refer to standard error of mean. Number of samples (biological and technical replicates) per species is indicated in parenthesis.

DOI: http://dx.doi.org/10.7554/eLife.19130.004

Figure 2.

Figure 2—figure supplement 1. Quality assessment of orthologs.

Figure 2—figure supplement 1.

(A) Percentage of ortholog sets filled up using consensus. Horizontal axis indicates the percentage of sequence length filled up by consensus. For example, 74% of the ortholog sets did not require filling up or were filled up <10% of the sequence length. Five percent of the ortholog sets were filled up 90–100% of the sequence length. (B) Standardized expression values of ortholog sets filled up using consensus. Within each ortholog set, the expression values were standardized to mean = 0 and standard deviation = 1. Distribution of the median pairwise DNA distance among (C) the Mouse Reference orthologs and (D) the expressed orthologs. The DNA distanced is based on Kimura 2- parameters distance.

The median nucleotide sequence identity for our ortholog sets with respect to mouse ranged from 83.2% (shrew) to 95.0% (African grass rat), and protein sequence identity from 88.0% (little brown bat) to 96.8% (African grass rat) (Figure 2B), consistent with the evolutionary distance of the species to mouse. The read alignment rates were also largely consistent across samples (Figure 2C). For a number of sequences with poor coverage, the consensus sequences of closely related species were used instead, but this did not significantly affect the results (Figure 2—figure supplement 1). After data filtering and normalization (Materials and methods), the expression of 9389 gene orthologs was reliably detected across the 28 samples (Supplementary file 1). For those species with publicly available genomes, ~10,000–11,000 genes could be reliably detected and the read counts also showed strong agreement (Pearson correlation coefficient 0.95–0.98 for log10 counts; Figure 1—source data 1C).

Gene expression patterns in fibroblasts follow phylogeny

To assess the gene expression patterns across the species, we performed Principal Component Analysis and projected the data on the first three Principal Components (Figure 3A). The samples segregated predominantly by their taxonomic relationship. For example, the species belonging to the sub-orders Sciuromorpha (chipmunk, red squirrel, and fox squirrel), Hystricomorpha (guinea pig, porcupine, and chinchilla), and Myomorpha (African grass rat, meadow vole, cotton rate, white-footed mouse, and deer mouse) separated clearly from one another (Figure 3A). The topology of the expression phylogram was also similar to the tree based on nucleotide sequences (Figure 3B), suggesting the expression patterns are influenced by phylogeny. In addition, the biological and technical replicates of the respective species clustered together, confirming that the within-species variation was generally smaller than the cross-species variation (Ma et al., 2015b).

Figure 3. Gene expression variation and correlation with longevity.

(A) Projection of the first three Principal Components (PCs) in Principal Component Analysis. Values in parenthesis indicate percentage of variance explained by each of the PCs. Points are colored by taxonomic order (same color scheme as in Figure 1) (B) Gene expression phylogram. Color of the nodes indicates the result of 1000 times bootstrap. (C) Overlap of genes associating with Adult Weight and indicated longevity traits. AW: Adult Weight; ML: Maximum Lifespan; FTM: Female Time to Maturity; MLres: Maximum Lifespan Residual; FTMres: Female Time to Maturity Residual. (D) Heat map showing expression patterns of the top enrichment pathways. Species are arranged in the order of increasing longevity (the four longevity traits are scaled between 0 and 1).

DOI: http://dx.doi.org/10.7554/eLife.19130.006

Figure 3.

Figure 3—figure supplement 1. Interaction network among the top hits in (A) positive and (B) negative correlation with longevity.

Figure 3—figure supplement 1.

The lines represent interaction based on STRING database (mouse genes). Selected gene names are colored based on the enriched pathways (see Table 1). Only the connected nodes are shown.

Expression of many genes correlates with longevity traits

To identify the genes with significant correlation to longevity, we performed regression by generalized least squares between the gene expression values and AW, as well as the four longevity traits (ML, FTM, MLres, and FTMres). The phylogenetic relationship of the species was incorporated in the variance-covariance matrix, and four different trait evolutionary models were tested to select the best models based on maximum likelihood (Materials and methods) (Lavin et al., 2008; Ma et al., 2015b). A two-step verification procedure was applied to assess robustness of the results (Ma et al., 2015b). Briefly, the potential outlier point was first identified and excluded to improve the regression fit (the regression slope p value was reported as ‘p value.robust’). Regression was then repeated by excluding each species, one at a time, to report the maximal (i.e. least significant) p value (‘p value.max’), to ensure the overall relationship did not depend on any single species.

We qualified as top hits those genes meeting both criteria of p value.robust < 0.01 (~11% FDR) and p value.max < 0.05. The numbers of top hits were 675 for AW, 812 for ML, 830 for FTM, 508 for MLres, and 793 for FTMres, with roughly equal proportions in positive and negative correlations (Table 1—source data 1A–F) and some overlap among the four longevity traits (Figure 3C). For most of the top hits, the directions of correlation were consistent across the four longevity traits (even for those that failed to reach statistical significance), suggesting there was a core set of longevity-associated genes and the minor inaccuracy in the reported lifespan data was unlikely to affect the overall results. On the other hand, the overlap with the hits identified by AW was much smaller (Figure 3C), indicating the observed correlations were not driven mainly by body mass differences. For the 827 top hits supported by two or more longevity traits, we performed pathway enrichment analysis using DAVID (Table 1, Table 1—source data 1G–H) (Huang da et al., 2009a, 2009b) based on mouse pathways.

Table 1.

Pathway enrichment analysis of genes with significant correlation with the longevity traits.

The genes were supported by at least two longevity traits (p value.robust < 0.01 and p value.max < 0.05). Pathway enrichment was performed using DAVID. The percentages of positive or negative correlating genes belonging to each pathway were indicated in parentheses. Only selected pathways are shown here. GO (BP): Gene Ontology (Biological Process). GO (BP): Gene Ontology (Molecular Functions). SP/PIR: SwissProt and Protein Information Resource. See Table 1—source data 1 for more details.

DOI: http://dx.doi.org/10.7554/eLife.19130.008

Table 1—source data 1. Phylogenetic regression of gene expression against longevity traits.
Regression against (A) Adult Weight; (B) Maximum Lifespan; (C) Female Time to Maturity; (D) Maximum Lifespan Residual; and (E) Female Time to Maturity Residual. ‘coef.all’, ‘p value.all’, and ‘q value.all’ refer to the regression slope, p value, and FDR-adjusted q value using all the species. ‘p value.robust’ and ‘q value.robust’ refer to the statistics after removing the potential outlier point. ‘p value.max’ and ‘q value.max’ refer to the maximal (least significant) regression p value and q value when each one of the species was left out, one at a time. Only genes with p value.robust<0.01 and p value.max<0.05 are shown. (F) Top hits identified by two or more longevity traits. The p value.robust against each of the four longevity traits (ML, FTM, MLres, and FTMres) as well as adult weight (AW) are shown. These genes were the input for pathway enrichment analysis. Pathway enrichment analysis of genes showing (G) positive and (H) negative correlation with longevity traits. Enrichment was performed using DAVID with default settings. Only the top 10 clusters are shown. (I) System level analyses of gene functions. The numbers of shared genes between longevity associated genes (either positively or negatively or both) and human aging genes, essential genes, transcription factor genes, and housekeeping genes are shown. The enrichment p value was calculated by Fisher’s exact test with two different background gene sets.
DOI: 10.7554/eLife.19130.009

Annotation cluster

Enriched terms and genes

No. of genes

p Value

Positive Correlation

Cluster No. 1

(15%)

GO (MF): adenyl nucleotide binding

50

5.25 × 10−3

GO (MF): nucleotide binding

64

1.21 × 10−2

Acly, Atad2, Atp2b4, Cdk2, Cdk20, Chd7, Chek1, Chkb, Cpsf7, D2hgdh, Dgkq, Dhx58, Dock6, Ero1lb, Etnk1, Fastkd5, Fn3krp, Gnai1, Guk1, Hk1, Hmgcr, Hnrnpd, Hyou1, Insr, Madd, Map4k5, Mastl, Mlkl, Mov10, Msh6, Mx2, Nadsyn1, Oplah, Pdk1, Pfkp, Phka2, Phkg2, Pkmyt1, Pms2, Pnkp, Ppp2r4, Prkar1b, Qrsl1, Rbm10, Rbm15b, Rbm38, Rhot2, Rnasel, Rps6ka2, Sacs, Sirt3, Slirp, Smarca1, Smarca5, Srsf9, Stk19, Stk36, Tbrg4, Tesk2, Thnsl1, Tia1, Top3a, Trpm4, Ttf2, Tyk2, Vps4a, Ythdc2

Positive Correlation

Cluster No. 2

(4%)

SP/PIR: DNA damage

14

1.16 × 10−3

SP/PIR: DNA repair

12

4.25 × 10−3

GO (BP): cellular response to stress

16

1.01 × 10−1

Bnip3, C17orf70, Chek1, Dtx3l, Ercc1, Errfi1, Fancg, Hif1a, Mapkbp1, Msh6, Myd88, Pms2, Pnkp, Prdx3, Prpf19, Pttg1, Rad51b, Rif1, Rnaseh1, Slx4, Tdp2, Terf1, Tinf2, Top3a, Wrap53

Positive Correlation

Cluster No. 4/5

(4%)

GO (BP): glucose metabolic process

11

1.22 × 10−3

GO (BP): hexose metabolic process

11

5.68 × 10−3

GO (BP): generation of precursor metabolites and energy

15

4.59 × 10−3

Aldh5a1, Atp2b4, Atp6v0d1, Atp6v0e2, Ero1lb, Fads1, Gbe1, Gpi1, Hexa, Hk1, Insr, Ndst1, Ndufa8, Pdk1, Pfkp, Pgp, Phka2, Phkb, Phkg2, Prkar1b, Sdhaf3, Tmx4, Tpi1, Trpm4, Tsc2

Positive Correlation

Cluster No. 6

(4%)

SP/PIR: chromatin regulator

11

1.61 × 10−2

GO (BP): chromosome organization

17

2.22 × 10−2

Bnip3, Cenph, Chd7, Dtx3l, Ercc1, H2afv, Hdac2, Jade1, Kdm5d, Kmt2c, Pttg1, Rcor1, Rrp8, Smarca1, Smarca5, Smyd3, Terf1, Tinf2, Wdr5, Wrap53

Negative Correlation

Cluster No. 1

(9%)

GO (BP): modification-dependent protein catabolic process

27

2.39 × 10−4

SP/PIR: ubiquitin conjugation pathway

26

3.35 × 10−4

GO (BP): proteolysis

36

1.09 × 10−2

Adamts2, Agtpbp1, Anapc4, Atg10, Atg4a, Atg7, Btbd1, Ctsl, Ctsz, Dcaf10, Dda1, Dpp8, Fbxl17, Fbxl20, Fbxo18, Fbxw2, Kcmf1, Map1lc3b, Med8, Mmp2, Mycbp2, Oma1, Pcsk5, Pgpep1, Pmepa1, Ppp2r5c, Rad18, Rfwd2, Rnf14, Rnf2, Rnf6, Sumo3, Tpp2, Ube2b, Ube2v1, Ufm1, Vhl

Negative Correlation

Cluster No. 2

(9%)

GO (BP): protein localization

38

4.67 × 10−5

GO (BP): protein transport

34

7.99 × 10−5

Agap1, Akap7, Ap3d1, Atg10, Atg4a, Atg7, Bax, Cav1, Clpx, Cnih1, Col4a3bp, Cry2, Dirc2, Ergic2, Fdx1l, Fkbp15, Gabarapl2, Gdi2, Gm10273, Golt1b, Hspa9, Ift46, Ipo4, Kif1bp, Kpna4, Laptm4a, Lrp4, mt-Nd4, Mtch1, Ndel1, Ndufb11, Necap1, Ppp3ca, Rab18, Rab2a, Rab6a, Rhot1, Sar1a, Sec22a, Sec31a, Sec62, Slc25a12, Slc29a1, Slc33a1, Slc35a4, Snx12, Snx13, Stx17, Timm8a1, Tomm6, Trappc6b, Trp53, Tsg101, Vps36, Vps53, Ywhag

Negative Correlation

Cluster No. 3

(18%)

GO (BP): regulation of transcription

74

1.62 × 10−5

SP/PIR: transcription regulation

55

1.04 × 10−3

Actl6a, Agtpbp1, Ak6, Anp32a, Anp32e, Atf6b, Bckdha, Bmi1, Ccdc59, Cd3eap, Cdc5l, Cggbp1, Clk2, Cnbp, Cops7a, Crtc3, Cry2, Csrp2, Ebna1bp2, Ehmt2, Elk4, Ergic2, Fbxo18, Fip1l1, Fosb, Foxo3, Gatad2b, Gid8, Gmcl1, Gtf2h1, Gtf2h2, Gtf2h5, Harbi1, Hlx, Hmga1-rs1, Hnrnpab, Hnrnpf, Ift57, Ing2, Ints4, Ipo4, Jund, Klf11, Klf2, Klf4, Klf9, Kpna4, Mafb, Mapk1, Mdm4, Med16, Med17, Med31, Med8, Mef2a, Mettl8, Mmp2, Mnt, Morf4l2, Mta1, Mtdh, Mxd1, Mycbp2, Nabp2, Ncor2, Neo1, Nfe2l2, Nr1d2, Papd4, Parp2, Phf12, Phlpp1, Pkig, Pomp, Pop5, Ppp1r8, Ppp2r5c, Ppp3ca, Ptbp1, R3hdm4, Rab18, Rad18, Rbbp4, Rfwd2, Rnf14, Rnf2, Rnf6, Rps6ka4, Rrs1, Sap30l, Sav1, Scoc, Sfmbt1, Sin3b, Snrk, Sqstm1, Srpk2, Ssbp1, Tep1, Tgfbr3, Trim35, Trip6, Trp53, Tsg101, Ube2b, Ube2v1, Ubtf, Ufm1, Vhl, Vps36, Wiz, Xrcc5, Yeats4, Zbtb14, Zfp414, Zfp637, Zfp655, Zfp710, Zfp821

Genes showing positive correlation with lifespan

The top pathways for the genes with positive correlation included ‘nucleotide binding’ (15% of the genes with positive correlation to longevity), ‘DNA repair’ (4%), ‘glucose metabolic process’ (4%), and ‘chromosome organization’ (4%) (Table 1, Figure 3D). The ‘DNA repair’ genes included those in DNA mismatch repair (Msh6, Pms2), nonhomologous end joining and possibly other repair pathways (Pnkp), nucleotide excision repair and DNA double-strand break repair (Ercc1), Fanconi anemia-associated DNA damage response network (C17orf70, Fancg), and protection of telomeres (Rif1, Terf1, Tinf2). The products of checkpoint kinase Chek1 and anaphase promoting complex substrate Pttg1 were regulators of cell cycle.

Among the other genes, Hif1a encodes the alpha subunit of hypoxia-inducible factor 1 (HIF-1), a key transcription factor in mediating the metabolic responses to hypoxia, whereas Prdx3 encodes mitochondrial peroxiredoxin that regulates redox homeostasis. In particular, Pnkp (Figure 4A), Prdx3, and Rif1 reached statistical significance in all four longevity traits (Table 1—source data 1F). Consistent with the findings, over-expression of hif-1 in C. elegans was shown to promote longevity (Zhang et al., 2009), whereas deletion of rif1 and msh6 in yeast (Austriaco and Guarente, 1997; Laschober et al., 2010), knockout of prdx3 in C. elegans (Ha et al., 2006), and disruption of Ercc1 in mouse (Weeda et al., 1997) were all detrimental and led to decreased lifespan. Several previous studies also suggested that long-lived species generally have enhanced DNA repair capacity (Cortopassi and Wang, 1996), higher poly (ADP-ribose) polymerase activity (Grube and Bürkle, 1992), up-regulation of genes in base-excision repair and superoxide metabolic process (Fushan et al., 2015), as well as reduced free-radical production (Perez-Campo et al., 1998), reduced oxidant generation (Sohal et al., 1995), and less oxidative damage to nuclear DNA (Adelman et al., 1988) and mitochondrial DNA (Barja and Herrero, 2000), although the degree of contribution toward the observed differences in lifespan varied and might be affected by several confounding effects (Debrabant et al., 2014; Montgomery et al., 2012; Promislow, 1994).

Figure 4. Selected genes and stress resistance conditions with significant correlation to longevity.

Figure 4.

(A) Pnkp and (B) Nadsyn1 show positive correlation with the longevity traits. (C) Trp53, (D) Bax, (E) Mapk1, and (F) Jund show negative correlation with the longevity traits. In each plot, the gene expression values (vertical axis) and the longevity traits (horizontal axis; FTM: Female Time to Maturity; FTMres: Female Time to Maturity Residual) are centered at 0 on log10 scale and then transformed by the best-fit variance-covariance matrix under phylogenetic regression (i.e. to remove the phylogenetic relationship). The potential outlier point has been removed and the remaining points are shown on the plot and colored by taxonomic group (same color scheme as in Figure 1). The regression slope p value (i.e. p value.robust) and R2 value are indicated. Error bars indicate standard error of mean. Resistance to (G) cadmium and (H) paraquat treatments. In each plot, the lethal dose (LD50) values (vertical axis) and the longevity traits (horizontal axis; ML: Maximum Lifespan) are plotted on ordinary scale (without log transformation). The regression slope p values are 9.16 × 10−3 and 1.39 × 10−2, respectively.

DOI: http://dx.doi.org/10.7554/eLife.19130.010

‘Glucose metabolic process’ included the gene products of hexokinase (Hk1), glucose phosphate isomerase (Gpi1), triose phosphate isomerase (Tpi1), phosphofructose kinase (Pfkp), and pyruvate dehydrogenase kinase (Pdk1), which are involved in glycolysis/gluconeogenesis. The glucan branching enzyme (encoded by Gbe1) and several phosphorylase kinases (encoded by Phka2, Phkb, Phkg2) regulate the metabolism of glycogen. In addition, the genes coding for NAD synthetase (Nadsyn1), which is involved in converting nicotinate adenine dinucleotide (NaAD) to nicotinamide adenine dinucleotide (NAD), also showed positive correlation with all four longevity traits (Figure 4B). Previously, it was observed that NAD+ levels declined with age and affected SIRT1 functions, whereas supplementation with NAD+ precursors reversed the aging phenotypes in mouse muscle (Gomes et al., 2013), and overexpression of SIRT1 in mouse brain could protect against aging-dependent circadian changes (Chang and Guarente, 2013). Calorie restriction also increases the NAD+/NADH ratio in yeast (Lin et al., 2004). As our study did not directly quantify the NAD+/NADH ratio, it remains to be seen if the high Nadsyn1 expression in the fibroblasts of the long-lived species affects these metabolites.

Genes showing negative correlation with lifespan

With regard to the top hits showing negative correlation, the major enriched pathways included ‘proteolysis’ (9% of the genes with negative correlation to longevity), ‘protein transport/localization’ (9%), and ‘regulation of transcription’ (18%) (Table 1, Figure 3D). For ‘proteolysis’, we observed relatively low expression of the genes coding for E2 ubiquitin-conjugating enzyme (Ube2b, Ube2v1), E3 ubiquitin-protein ligase (Rad18, Mycbp2), ubiquitin-like modifier (Sumo3, Ufm1), as well as several proteins containing RING finger domain (Rnf2, Rnf6, Rnf14, Rfwd2) or F-box domain (Fbxl17, Fbxl20, Fbxo18, Fbxw2), both of which are known to be involved in the ubiquitination pathway. Also, low expression was observed for the genes encoding autophagy related proteins (Atg4a, Atg7, Atg10) and lysosomal cysteine proteinases (Ctsl, Ctsz). The genes implicated in ‘protein transport/localization’ included several vesicle trafficking proteins (Sec22a, Sec31a, Sec62, Golt1b), mitochondrial membrane translocases (Timm8a1, Tomm6), and nuclear transport receptors (Ipo4, Kpna4). As for ‘regulation of transcription’, we observed down-regulation of the genes coding for mediator complex subunits (Med8, Med16, Med17, Med31), zinc finger proteins (Zfp414, Zfp655, Zfp637, Zfp710, Zfp821), Kruppel-like factors (Klf2, Klf4, Klf9, Klf11), and members of the MYC/MAX/MAD network of transcription factors (Mxd1, Mnt).

Interestingly, the pathways related to ‘response to DNA damage’ and ‘cellular response to stress’ were also enriched (5% of the genes with negative correlation to longevity). A closer examination revealed that the enrichment signal was due to a number of genes involved in apoptosis regulation, including the tumor suppressor TP53 (encoded by Trp53), BCL-2 associated X protein BAX (encoded by Bax), transcription factor FOXO3 (encoded by Foxo3), as well as mitogen-activated protein (MAP) kinase (encoded by Mapk1) (Figure 4C–E); they were therefore distinct from those genes directly involved in DNA repair (and found above to have positive correlation with longevity). Several other growth signaling factors, such as transforming growth factor beta (TGF-β) receptor (encoded by Tgfbr3) and transcription factor JunD (encoded by Jund), were also relatively low in longer lived species (Figure 4F). In particular, the transcription factor FOXO3 can be activated by oxidative stress (Essers et al., 2004), and the genetic variation within the FOXO3A gene was found to be strongly associated with human longevity (Willcox et al., 2008).

Genes enriched in network interaction and housekeeping functions

To understand the regulatory network among the top hits, we visualized the protein-protein interaction using the STRING database (Jensen et al., 2009). The results revealed significant network interaction among the genes with positive correlation and those with negative correlation (p value < 10−10 in both cases; Figure 3—figure supplement 1), suggesting that the longevity-correlating genes were indeed functionally related. Next, we analyzed the system level functions of the top hits to determine if they belonged to the known categories of ‘Aging genes', 'Essential genes’, 'Housekeeping genes’ or ‘Transcription Factor genes’ (Table 1—source data 1I). Interestingly, close to 40% of the top hits belonged to the ‘Housekeeping genes’ (Fisher exact test p value = 3.646×10−26), whereas the other categories were much less significant (Table 1—source data 1I). Therefore, the longevity variation across these species was accompanied by the coordinated modulation of a large number of genes with housekeeping functions in a systemic manner.

Fibroblast resistance to lethal stresses and toxicity

The longevity-associated expression patterns identified above suggested that the longer lived species might be more efficient at handling and repairing cellular damage. It was previously demonstrated that skin-derived fibroblasts from long-lived rodent species were more tolerant of the stress conditions induced by cadmium, hydrogen peroxide, heat, and DNA alkylating agent methyl methanesulfonate (MMS), and were more resistant to the metabolic inhibition by rotenone treatment and in low-glucose medium (Harper et al., 2007). To see if the same effects could be observed in our study, we subjected the primary fibroblasts to six different stress conditions: treatments with cadmium, hydrogen peroxide, MMS, paraquat, and thapsigargin (inhibitor of sarco/endoplasmic reticulum Ca2+ ATPase), as well as to low-glucose culture medium. As expected, the results showed positive correlation between ML and the resistance to cadmium and paraquat (Figure 4G,H), although the other conditions did not reach the statistical threshold of p value < 0.05.

Metabolites correlating with longevity traits

For 12 of the rodent species, we also performed metabolic analyses (Townsend et al., 2013) (Figure 1—source data 1D). After data filtering and normalization, 144 water-soluble metabolites and 82 lipids were reliably detected across the 22 biological samples (Supplementary file 2). Principal Component Analysis (Figure 5A) and the phylogram based on metabolite levels (Figure 5B) both indicated that the metabolic profiles of these species, like the gene expression, segregated according to phylogeny, although the patterns were less clear-cut than those based on the RNAseq dataset. This might be partly due to the much smaller number of metabolites detected compared to the genes (226 metabolites vs. 9389 genes). Nevertheless, the biological and technical replicates clustered together (Figure 5B), suggesting that the within-species variation was relatively small.

Figure 5. Metabolite variation and correlation with longevity.

(A) Projection of the first three Principal Components (PCs) in Principal Component Analysis. Values in parenthesis indicate percent of variance explained by each of the PCs. Points are colored by taxonomic order (same color scheme as in Figure 1) (B) Metabolite phylogram. Color of the nodes indicates the result of 1000 times bootstrap. (C) Overlap of metabolites associating with Adult Weight and longevity traits. AW: Adult Weight; ML: Maximum Lifespan; FTM: Female Time to Maturity; MLres: Maximum Lifespan Residual; FTMres: Female Time to Maturity Residual. (D) Amino acids showing positive correlation with Maximum Lifespan (ML). In each plot, the amino acid levels (vertical axis) and the longevity traits (horizontal axis) are centered at 0 on log10 scale and then transformed by the best-fit variance-covariance matrix under phylogenetic regression (i.e. to remove the phylogenetic relationship). The potential outlier point has been removed and the remaining points are shown on the plot and colored by taxonomic group. The regression slope p value (i.e. p value.robust) and R2 value are indicated. Error bars indicate standard error of mean.

DOI: http://dx.doi.org/10.7554/eLife.19130.011

Figure 5.

Figure 5—figure supplement 1. Amino acid levels in primate and bird fibroblasts correlate positively with species maximum lifespan.

Figure 5—figure supplement 1.

Each point represents a different species of bird (green triangles) or non-human primate (orange circles), with linear regression lines shown separately for each group of species. Data for human fibroblasts are presented (orange triangle; 'H'), but did not contribute to the regression lines or significance and slope estimates shown in Table 2.

To identify the metabolites with significant correlation with the longevity traits, we also applied the phylogenetic regression method described above. At the cut-off of p value.robust < 0.01 (~11% FDR) and p value.max < 0.05, 13 metabolites showed significant correlation with AW, 26 metabolites with ML, 20 metabolites with FTM, 16 metabolites with MLres and 19 metabolites with FTMres (Figure 5C, Table 2—source data 1A–F). Twenty-three of these metabolites were supported by two or more longevity traits. Pathway analysis revealed the enrichment of ‘common amino acids’ among the top hits with positive correlation, and ‘glycerophospholipids’ among the top hits with negative correlation (Table 2—source data 1G–H). In particular, several showed positive correlation with multiple longevity traits (Figure 5D, Table 2); so did a number of nucleotides/nucleosides including ADP, GDP, and adenosine. In terms of negative correlation, a number of lysophosphatidylchonline (LPC; e.g. C16:0 LPC, C18:0 LPC, C18:1 LPC) and lysophosphatidylethanolamine (LPE; e.g. C20:4 LPE, C22:6 LPE) showed significant relationship (Table 2—source data 1), which were consistent with the previous report of low LPC and LPE in long-lived mammals (Ma et al., 2015b). LPC levels were also previously reported to decrease with age but maintained in mice under caloric restriction (De Guzman et al., 2013).

Table 2.

Amino acid levels showing consistent positive correlation with longevity traits.

For the mammalian fibroblast dataset, the number of longevity traits (out of Maximum Lifespan; Female Time to Maturity; Maximum Lifespan Residual; and Female Time to Maturity Residual) with significant positive correlation with the amino acid levels at two different cut-offs (p value.robust < 0.01 and p value.robust < 0.05) are shown. For the primate and bird fibroblast dataset, the regression was performed using primate data only, bird data only, and the pooled data of both. The regression slope p value < 0.05 are in bold.

DOI: http://dx.doi.org/10.7554/eLife.19130.013

Table 2—source data 1. Phylogenetic regression of metabolite levels against longevity traits.
Regression against (A) Adult Weight; (B) Maximum Lifespan; (C) Female Time to Maturity; (D) Maximum Lifespan Residual; and (E) Female Time to Maturity Residual. ‘coef.all’, ‘p value.all’, and ‘q value.all’ refer to the regression slope, p value, and FDR-adjusted q value using all the species. ‘p value.robust’ and ‘q value.robust’ refer to the statistics after removing the potential outlier point. ‘p value.max’ and ‘q value.max’ refer to the maximal (least significant) regression p value and q value when each one of the species was left out, one at a time. Only genes with p value.robust < 0.01 and p value.max < 0.05 are shown. (F) Top hits identified by two or more longevity traits. The p value.robust against each of the four longevity traits (ML, FTM, MLres, and FTMres) as well as adult weight (AW) are shown. These metabolites were the input for pathway enrichment analysis. Pathway enrichment analysis of metabolites showing (G) positive and (H) negative correlation with longevity traits. Enrichment was performed based on hypergeometric statistics. (I) Top hits identified by two or more longevity traits, using cut-off of p value.robust < 0.05. The p value.robust against each of the four longevity traits (ML, FTM, MLres, and FTMres) as well as adult weight (AW) are shown.
DOI: 10.7554/eLife.19130.014

Amino acid

Mammalian fibroblasts

Primate and bird fibroblasts

No. of longevity traits (out of four) with significant correlation

Regression slope p value with species maximum lifespan

Regression slope p value with species maximum lifespan residual

p value.robust < 0.01

p value.robust < 0.05

Primates only

Birds only

Primates and birds

Primates only

Birds only

Primates and birds

arginine

3

4

3.4 × 10−2

8.6 × 10−2

3.1 × 10−2

3.8 × 10−1

1.1 × 10−2

2.1 × 10−2

glutamate

2

4

6.5 × 10−2

1.8 × 10−2

1.1 × 10−2

4.6 × 10−2

2.8 × 10−1

1.3 × 10−1

histidine

0

4

9.4 × 10−2

6.0 × 10−2

4.3 × 10−2

2.3 × 10−1

1.4 × 10−1

1.7 × 10−1

leucine

2

4

2.9 × 10−3

6.0 × 10−2

4.8 × 10−3

1.4 × 10−2

5.9 × 10−1

2.3 × 10−1

lysine

3

3

9.8 × 10−3

8.2 × 10−2

1.4 × 10−2

9.1 × 10−2

2.9 × 10−1

2.5 × 10−1

methionine

1

3

3.2 × 10−1

1.4 × 10−2

2.7 × 10−2

3.0 × 10−1

3.0 × 10−2

4.9 × 10−2

phenylalanine

1

4

9.8 × 10−3

1.2 × 10−3

2.1 × 10−4

8.2 × 10−2

1.3 × 10−1

1.2 × 10−1

proline

1

4

4.4 × 10−3

3.9 × 10−4

3.6 × 10−5

3.5 × 10−2

1.2 × 10−1

5.4 × 10−2

tryptophan

2

4

9.2 × 10−3

7.8 × 10−4

1.2 × 10−4

2.6 × 10−2

2.5 × 10−1

1.5 × 10−1

tyrosine

1

3

3.2 × 10−1

8.8 × 10−3

1.8 × 10−2

4.3 × 10−1

1.7 × 10−1

2.9 × 10−1

valine

0

3

1.2 × 10−2

5.4 × 10−3

1.0 × 10−3

2.0 × 10−1

2.8 × 10−1

3.2 × 10−1

Validation of amino acid patterns in primate and bird fibroblasts

To further examine our observation of the positive correlation between amino acids and the longevity traits, we independently obtained and quantified the amino acid levels in a larger collection of primary fibroblasts from 15 primate species and 33 bird species. All 10 of the amino acids associated with lifespan in rodent fibroblasts (arginine, glutamate, histidine, leucine, lysine, methionine, phenylalanine, proline, tryptophan, tyrosine, and valine) were also found to have a significant positive association with lifespan in bird and primate fibroblasts (Table 2; Figure 5—figure supplement 1). The associations were particularly strong for amino acids with hydrophobic side chains. When we adjusted for the effects of body mass, the observed relationships weakened significantly (Table 2), likely due to the strong correlation between AW and ML. Nevertheless, the same weakening was also evident in the rodent fibroblasts (Table 2—source data 1), indicative of the consistency in the trends. Overall, the positive correlation between fibroblast amino acid levels and species longevity was a feature consistent across rodents, primates, and birds, indicating that some of the longevity signatures identified here may be representative of and generalized to other species.

Discussion

All lines of mammals descended from the same common ancestor over the previous 230 million years and have since undergone remarkable diversification in body size, metabolic rate, fertility, and longevity, with corresponding changes in the gene expression and metabolite landscape (Fushan et al., 2015; Ma et al., 2015b). As fibroblasts can be obtained without sacrificing animals and can be cultured under standardized conditions, it is of great interest to determine if their gene expression and metabolite patterns represent lifespan variation across mammals. Fibroblasts are also amenable to experimental manipulation. On the other hand, cross-species gene expression analyses are often hampered by the lack of publicly available genomes and gene orthology information, especially for those species not commonly studied. Using primary fibroblasts from 16 species of rodents, bats, and shrew, we developed a pipeline for generating species-specific ortholog sets, profiled gene expression by RNAseq and metabolites by mass spectrometry, and identified the molecular features associated with longevity traits.

Our pipeline can be easily extended for a larger number of species. We defined gene orthology based on reciprocal best hit in BLAST (Tatusov et al., 1997) and ignored the issues of gene duplication and gene loss. Sequence fragments and missing sequences were filled up using consensus data from the other species. While these steps unavoidably introduced inaccuracy within our species-specific ortholog sequences, they did not affect the overall read alignment results (Figure 2B–C, Figure 2—figure supplement 1). Furthermore, we observed no significant differences in sequence divergence between those ortholog sets showing correlation to longevity and those that did not (Wilcoxon Rank Sum Test p value = 0.32; Figure 2—figure supplement 1D), so the degree of longevity correlation was not biased by the degree of sequence conservation.

The gene expression findings revealed a clear segregation based on phylogeny (Figure 3A–B), suggesting that evolutionary relationships significantly influenced the expression patterns. On the other hand, the metabolite patterns were less clear-cut (Figure 5A–B), which might be attributed to the smaller number of species and metabolites. Using phylogenetic regression and a two-step verification procedure, we identified a list of genes and metabolites with significant correlations to multiple longevity traits. The pathways of ‘nucleotide binding', 'DNA repair’, 'chromosome organization’, and ‘glucose metabolic process’ were enriched among the genes with positive correlation with longevity, whereas ‘proteolysis’, ‘protein transportation/localization’ and ‘regulation of transcription’ were enriched for genes showing a negative correlation. Furthermore, a significant number of these longevity-correlating genes are involved in ‘housekeeping’ functions, implying that lifespan variation across species is often accompanied by coordinated shifts in the gene expression landscape and modulation of fundamental biological processes.

The link between proteolysis/autophagy and aging has been proposed by a number of authors. In general, proteolytic functions decline and oxidized proteins increase with age, and autophagy genes are required for the lifespan extension effects of Insulin/IGF-1 signaling and dietary restriction (Chondrogianni and Gonos, 2008; Hansen et al., 2008; Kenyon, 2010; Kevei and Hoppe, 2014; Löw, 2011; Meléndez et al., 2003; Rubinsztein et al., 2011; Starke-Reed and Oliver, 1989; Vernace et al., 2007). Activation of proteasome or autophagy has been shown to extend lifespan in C. elegans (Chondrogianni et al., 2015; Ghazi et al., 2007), yeast (Kruegel et al., 2011), and flies (Simonsen et al., 2008). Immunoproteasome and proteasome activity was also elevated in the livers of long-lived Snell dwarf mice and in mice exposed to drugs known to extend lifespan (Pickering et al., 2015). On the other hand, our results suggest that fibroblasts from the longer lived animals actually have lower expression levels of genes involved in proteolysis, autophagy, and apoptosis but higher expression of genes related to DNA repair and maintenance. In particular, the genes coding for the tumor suppressor TP53, apoptosis regulator BAX, and several growth and proliferation signaling pathways were all down-regulated in the fibroblasts of the longer lived species (Figure 4C–F). One possible interpretation may be that the longer lived species generate less damage and/or have better repair mechanisms, so that the cells rely less on proteolysis, autophagy and apoptosis. Previous studies reported enhanced DNA repair capacity and reduced oxidative damage in longer lived species (Adelman et al., 1988; Cortopassi and Wang, 1996; Grube and Bürkle, 1992; Perez-Campo et al., 1998; Sohal et al., 1995). Down-regulation of the ubiquitin ligase complex was also reported in the liver of longer lived mammalian species (Fushan et al., 2015). In agreement, our toxicology experiments confirmed that the fibroblasts of longer lived species were more resistant to oxidative stress induced by cadmium and paraquat treatments. In terms of metabolites, the pattern of low LPC and LPE among long-lived species was consistent with previous reports, and the positive correlation between amino acids and longevity was independently validated using fibroblasts from multiple species of primates and birds, suggesting that these changes in cell biology are likely to have evolved, independently, in each of these separate lineages in association with slower aging and longer lifespan.

On the other hand, several possible caveats in our data warrant additional attention. The levels of gene expression could be influenced by confounding factors such as gene length and proximity to other genes (Chiaromonte et al., 2003), and our definition of orthology might have missed out those genes with less conserved sequences. Our analyses were limited to those genes expressed in fibroblasts, and the effects of different spliced isoforms were not captured by our data. Furthermore, although our RNA sequencing and metabolic measurements were performed on cells of second or third passage, it was nevertheless possible that in vitro culture conditions might have introduced changes in chromatin architecture (Zhu et al., 2013), or that differences in stress-related pathways might reflect differential responses to the culture conditions. In addition, the use of ML as an aging research metric can be problematic for the less well documented animals due to the small sample size, high variance, and reliance on a single individual per species (Moorad et al., 2012). Although FTM is less prone to reporting bias and shows strong correlation with ML (Spearman correlation coefficient 0.87), it may be influenced by seasonality factors. Other parameters with better statistical properties, such as mean adult lifespan and 90th quantile of longevity (Moorad et al., 2012), should be used in place of ML in the long run, although at the moment such records are available for only a limited number of species.

Overall, our study supports the idea that gene expression, and to some degree metabolite levels, in fibroblast cultures can uncover differences in cell biology and metabolism that correspond to longer life. Apparently, these expression patterns are preserved when the intraorganismal environment is removed and cells instead are subjected to standardized cell culture conditions in the lab setting. This makes fibroblasts a particularly attractive experimental system to examine and manipulate molecular patterns, with gene expression (or metabolite patterns) as a readout. While our study represents an initial study, this approach can be extended to a larger group of species and samples, refining the molecular signatures and then manipulating them via genetic and environmental manipulations. Ultimately, this should reveal the genetic basis for differences in species longevity and lead to new strategies for targeting them, thereby shifting cells, and ultimately organisms, to the state of cells from related longer-lived species.

Materials and methods

Sample collection

Primary skin fibroblast samples were collected from shrew (Blarina brevicauda), big brown bat (Eptesicus fuscus), little brown bat (Myotis lucifugus), guinea pig (Cavia porcellus), porcupine (Erethizon dorsatum), chinchilla (Chinchilla lanigera), chipmunk (Tamias striatus), fox squirrel (Sciurus niger), red squirrel (Sciurus vulgaris), beaver (Castor canadensis), gerbil (Meriones unguiculatus), African grass rat (Arvicanthis niloticus), meadow vole (Microtus pennsylvanicus), cotton rat (Sigmodon hispidus), white-footed mouse (Peromyscus leucopus), and deer mouse (Peromyscus maniculatus brandii) (Figure 1—source data 1). The post-pubertal animals (gender and ages were not recorded) were caught opportunistically in an area extending approximately 400 km north and 80 km south of Ann Arbor, MI, USA (Harper et al., 2007). Abdominal skin areas were sterilized with 70% ethanol wipes and biopsies of at least 5 mm by 5 mm in area were obtained and placed in complete media (CM) made of Dulbecco's modified Eagle medium (DMEM, high-glucose variant, Gibco-Invitrogen, Carlsbad, CA) supplemented with 20% heat-inactivated fetal bovine serum, antibiotics (100 U mL−1 penicillin and 100 µg mL−1 streptomycin; Sigma, St. Louis, MO) and 0.25 µg mL−1 of fungizone (Biowhittaker-Cambrex Life Sciences, Walkersville, MD) on ice and shipped overnight to our laboratory (Harper et al., 2007). Biological replicates (i.e. tissues from different individuals) and technical replicates were collected on selected species (Figure 1—source data 1).

Cell culture

The conditions for establishment and maintenance of the cultures have been reported previously (Harper et al., 2007; Murakami et al., 2003; Salmon et al., 2005). Briefly, trypsinized cells were grown to 90% confluence, and we found no significant differences among species in the interval between initial seeding and initial confluence (Harper et al., 2007). Cells were then harvested and placed in a new culture flask, fed at days 3 and 7, and then subcultured to a fixed density of 7.5×105 cells for 75 cm2 flask. These cells were then harvested 7 days later and cryopreserved at 106 cells per vial.

Production of cells for RNA sequencing and metabolite profiling always started by thawing a vial of cryopreserved cells and allowing them to expand until the culture had produced sufficient cells (at least 30 × 106 cells) for analysis. These cultures were kept under low-oxygen conditions (3% O2) after thawing to minimize selection for resistance to O2 toxicity (Busuttil et al., 2003; Parrinello et al., 2003). Cells were harvested using trypsin and pelleted by centrifuging for 5 min at 230 rcf. After counting, the cells were divided into aliquots of 10 × 106 cells. Two washes with PBS (-Ca,-Mg) were performed, any excess PBS was drained, and the pellets were frozen at −80°C. Technical replicates were made by growing a minimum of 60 × 106 cells and labeling half of the cells after counting as separate samples.

Life history data of the species

The Adult Weight (AW), Maximum Lifespan (ML) and Female Time to Maturity (FTM) data of the species (or if not available, for a closely related species) were obtained from the Animal Ageing and Longevity (AnAge, RRID:SCR_001470) Database (Tacutu et al., 2013). In addition, since both ML and FTM increase with AW, we calculated the body mass adjusted residuals (i.e. MLres and FTMres), to represent the ratio between the observed longevity and the expected longevity based on body mass (Ma et al., 2015b; Tacutu et al., 2013). Two allometric equations were used to calculate the residuals. The MLres equation, MLres = ML/(4.88×AW0.153), was based directly on the documentation of the AnAge database (http://genomics.senescence.info/help.html#anage). The FTMres equation, FTMres = FTM/(78.1×AW0.217), was based on linear regression using the FTM and body mass records of 1330 mammalian species in the AnAge database.

RNA sequencing

RNAseq libraries were prepared as previously described (Fushan et al., 2015). Paired end sequencing was done on the Illumina HiSeq2000 platform generating approximately 30 to 75 million reads per sample, with read length 50 or 100 nucleotides (Figure 1—source data 1). The raw data were processed by Cutadapt (RRID:SCR_011841) (Martin, 2011) to remove low-quality reads.

Species-specific ortholog sets and expression values

Reference genomes were publicly available for five species (Eptesicus fuscus, Myotis lucifugus, Cavia porcellus, Chinchilla lanigera, Peromyscus maniculatus brandii). To ensure consistency across the entire dataset, we developed the following pipeline to identify species-specific ortholog sets, map the reads and obtain expression values (Figure 2).

Step 1: generate mouse reference. Based on the Mus musculus Ensembl genome and annotation (release 78) (RRID:SCR_006773), the longest transcript was extracted for each protein-coding gene locus, after confirming the presence of start and stop codons and the proper reading frame. Those transcripts containing highly repetitive or highly similar sequences were identified and removed using BLAST (RRID:SCR_004870) (at e-value cut-off 10−6) (Camacho et al., 2009). This generated the Mouse Reference, representing the coding sequences of 16,816 unique protein-coding genes (Supplementary file 1).

Step 2: identify species-specific ortholog sets. For each species, the transcriptome was assembled de novo from the RNAseq reads using Trinity (RRID:SCR_013048) (Grabherr et al., 2011). BLAST (with ‘dc-megablast’ option) was performed between Mouse Reference and the assembled transcriptome (and the published genome, if available) of each species to identify the reciprocal best hits (Tatusov et al., 1997). The sequences were trimmed down to open reading frame (i.e. flanked by start and stop codons) using Exonerate (Slater and Birney, 2005). Within each ortholog sets, multiple sequence alignment was performed using MUSCLE (RRID:SCR_011812) (Edgar, 2004) and the percentage of sequence identity was assessed by MView (Brown et al., 1998). For the sequence fragments or missing sequences due to poor coverage, they were filled up using the consensus. We confirmed the filling up procedure did not significantly affect the read counting results (Figure 2—figure supplement 1). Seventy-four percent of the ortholog sets did not require filling up or were filled up <10% of the sequence length, whereas 5% of the ortholog sets were filled up 90–100% of the sequence length (Figure 2—figure supplement 1A). When the expression values were standardized to mean = 0 and standard deviation = 1 within each ortholog set, there was no significant bias against those ortholog sets with high percentage of filling up (Figure 2—figure supplement 1B).

Step 3: read mapping, counting, filtering and normalization. The RNAseq reads were mapped to the species-specific ortholog sets using STAR (Dobin et al., 2013), with an average read alignment rate of ~40% (Figure 1—source data 1). As comparison, read mapping to publically available genomes achieved an average alignment rate of ~85% (Figure 1—source data 1). The lower alignment rate to the species-specific ortholog sets was likely due to the exclusion of 5’ and 3’ untranslated regions, repetitive or highly similar sequences, and introns. Nevertheless, the alignment rates were largely similar across the samples and species (Figure 2C). Read counting was performed by featureCounts (RRID:SCR_012919) (Liao et al., 2014) and those ortholog sets with too high counts (i.e. read counts contributing to >5% of the total counts; three orthologs were removed this way) or too low counts (i.e. <10 counts in four or more samples) were discarded. The library sizes were then scaled by trimmed mean of M-values (TMM) method, log10-transformed, and quantile-normalized (Robinson and Oshlack, 2010). The final expression set consisted of 9389 gene orthologs across 28 samples. Shapiro Test confirmed normalcy assumption was valid for 89% of the genes on log10 scale. The pairwise DNA distance within each ortholog set was calculated based on the Kimura 2- parameters distance (Kimura, 1980).

Metabolite profiling and data processing

For rodent cells, the metabolite levels were quantified by mass spectrometry as previously described (Townsend et al., 2013). From the raw metabolite measurements, we only kept metabolites with <10% missing values. The raw values were normalized separately for the three collection modes (water soluble positive ionization mode ‘HILIC-pos’, water soluble negative ionization mode ‘HILIC-neg’, and lipid mode ‘C8-pos’), first by the internal standards, and then by the total signals within each mode. The data were then log10-transformed and quantile normalized. The final expression set consisted of 226 metabolites across 22 samples. Shapiro Test confirmed normalcy assumption was valid for 90% of the metabolites on log10 scale.

Principal component analysis and phylograms

Principal component analysis was performed on the standardized expression values or metabolite values and the first three Principal Components were extracted. The phylograms were constructed using the neighbor joining method (Saitou and Nei, 1987), based on the distance matrix of 1 minus Pearson correlation coefficient of the standardized expression values or metabolite values. The reliability of the branching patterns was assessed by 1000 times bootstrap.

Phylogenetic regression and two-step verification procedure

To identify genes or metabolites with significant correlation to the longevity traits, regression was performed using the generalized least square approach, by incorporating the phylogenetic relationship in the variance-covariance matrix (Felsenstein, 1985; Ma et al., 2015b; Revell, 2012). As previously described (Ma et al., 2015b), four different trait evolution models (‘null’, ‘Brownian motion', 'Pagel’s lambda’, and 'Ornstein-Uhlenbeck’) were tested and the best fit model was selected based on maximum likelihood.

A two-step procedure was applied to verify the robustness of the results (Ma et al., 2015b). In the first step, the species whose exclusion would lead to most improvement in the slope p value (i.e. a potential outlier), was identified and removed. The regression p value of this step was reported as ‘p value.robust’. In the second step, each of the remaining species was removed, one at a time, and the regression was repeated. The largest (i.e. least significant) p value of this step was reported as ‘p value.max’. The False Discovery Rate adjusted values were ‘q value.robust’ and ‘q value.max’, respectively. To qualify as a top hit, we required a gene to have p value.robust < 0.01 and p value.max < 0.05. For pathway enrichment purposes, we further required that the genes were identified as a top hit in two or more longevity traits (ML, FTM, MLres or FTMres).

Pathway enrichment analysis and interaction network

For the genes, pathway enrichment analysis was performed using DAVID (RRID:SCR_003033) (Huang da et al., 2009a, 2009b). The 16,816 unique protein-coding genes in Mouse Reference were set as the background and the pathways were based on Mus musculus. For those genes showing positive and negative correlation with longevity (supported by two or more longevity traits), we queried Gene Ontology (‘GO Term’; Biological Process and Molecular Functions only), SwissProt and Protein Information Resource (‘SP PIR Keywords’), and Kyoto Encyclopedia of Genes and Genomes (‘KEGG Pathway’). For comparison, pathway enrichment was also performed using only the 9389 expressed orthologs as background. STRING (RRID:SCR_005223) version 10 (Jensen et al., 2009) was used to visualize the interaction network among the top hits, based on the mouse genes. The required score was set as 400 and the network output was generated using R package (RRID:SCR_006442) ‘STRINGdb’. Selected nodes were highlighted based on the enriched pathways.

We also analyzed the association between longevity-associated genes (either positively or negatively or both) and each of four functional groups of genes – aging genes, essential genes, transcription factor genes, and housekeeping genes. These human gene sets were originally collected and analyzed in a previous study (Zhang et al., 2016). Human aging genes were obtained from GenAge (RRID:SCR_010223) (build 17) (de Magalhães and Toussaint, 2004). They include both genes related to fundamental human aging processes and those associated with human longevity based on evidence from human and model organisms. Human essential genes are the human orthologs of mouse essential genes, whose deletions result in prenatal, prenatal or postnatal lethality in mouse. Human transcript factor genes were from Panther database (RRID:SCR_004869) (Mi et al., 2013). Human housekeeping genes were from (Eisenberg and Levanon, 2013). Housekeeping genes are considered to be involved in basic cell maintenance, and thus expected to maintain relatively constant expression levels in different cells and conditions (Eisenberg and Levanon, 2013). For the enrichment analysis, we used two different sets of background genes. One includes all the orthologs tested (i.e. 16,816 genes), and the other only genes expressed in fibroblasts (i.e. 9389 genes). All mouse genes were mapped to their human orthologs through Ensembl BioMart (RRID:SCR_002344) (Smedley et al., 2015), and only genes with one to one mapping relationship were used. In this analysis, we used 14,749 and 8809 human orthologs as background genes. Enrichment statistics were based on Fisher’s exact test.

For the metabolites, pathway information was obtained from ConsensusPathDB (RRID:SCR_002231) (Kamburov et al., 2009) and Human Metabolome Database (HMDB) (RRID:SCR_007712) (Wishart et al., 2013). For ConsensusPathDB, only pathways with known KEGG IDs were incorporated. Analysis was performed on pathways with at least 5 but less than 100 metabolites. Enrichment statistics was based on a hypergeometric distribution (Tavazoie et al., 1999). Odd ratios and expected counts were calculated as previously described (Gentleman et al., 2013).

Evaluation of amino acid levels in bird and primate fibroblast cell lines

An untargeted metabolomics screen was conducted using fibroblasts from 32 bird species and 13 species of non-human primates. The detailed methods were described in McDonnell et al. (2013). Briefly, frozen cell cultures were homogenized in water, volumes were adjusted based on protein concentration in the extract, and proteins were precipitated by ethanol containing recovery standards. Each extract was split into two aliquots (for LC-MS and GC-MS), dried down. LC-MS aliquot was re-suspended in water containing injection standards and analyzed on an Agilent 1200 LC/6530 qTOF LC-MS system using a Waters Acquity HSS T3 C18 column. GC-MS aliquot was derivatized by BSTFA and analyzed on Agilent 7890 A-5975C inert XL MSD GCMS instrument using HP-5MS 5% phenyl-methyl siloxane column (30m x 250 μm x 0.25 μm). Data extraction and analysis was performed using Agilent MassHunter Qualitative Analysis software and in-house metabolite libraries.

Although the original dataset contained information on 4383 metabolites, including 456 of known chemical identity, the analysis for this paper was restricted to the ten amino acids for which p<0.05 for association with maximum lifespan in the analysis of mammalian (i.e. rodent, shrew and bat) fibroblasts (Figure 5 and Table 2—source data 1). A regression analysis was performed using a model in which the dependent variable was the logarithm of the species maximal lifespan, and the independent variables were the metabolite level and a categorical variable reflecting whether the species was bird or primate. The procedure tested the association between metabolite and lifespan in the entire set of species, but did not make the assumption that the slope was necessarily the same for birds as for primates. The resulting F-statistic was evaluated for significance based upon an empirically generated set of null distributions (for each metabolite) by permutation. When two or more of the untargeted features were annotated as corresponding to the same amino acid (seven cases), we tabulated the degree of association from the feature most strongly correlated with lifespan among the species studied. Although these would introduce some bias in favor of correlation, we noted that in four out of the seven cases, the multiple features were all significant at p value < 0.05; in the other three cases, the features were all significant at p value < 0.2. Parallel analyses were also done for bird species and for non-human primate species independently.

Resistance of rodent fibroblasts to lethal and metabolic stresses

The methods used are as previously described (Harper et al., 2007; Murakami et al., 2003; Salmon et al., 2005). Briefly, each test procedure began by culturing the cells at a density of 3 × 104 cells in 100 µL CM in 96-well microtiter plates for 24 hr, followed by a period of 24 hr in medium lacking serum but containing 2% bovine serum albumin (BSA, Sigma) with antibiotics and fungizone at the same concentration as CM. For assessment of resistance to H2O2, paraquat, and cadmium (Sigma), the cells in the 96-well plates were washed and exposed to the stress agent for 6 hr. For assessment of resistance to methyl methanesulfonate (MMS), the cells were incubated with MMS in DMEM for 24 hr, washed and then incubated with DMEM supplemented with 2% BSA, antibiotics, and fungizone for 18 hr. For assessment of cell metabolism in low-glucose medium, cells were incubated in DMEM containing a range of glucose concentrations for 1 hr. Survival was assessed by WST-1 tests. All incubations were at 37°C in a humidified incubator with 5% CO2 in air.

For calculation of the resistance of each cell line to chemical stressors, at each dose of chemical stressor, mean survival was calculated for triplicate wells for each cell line. The LD50, i.e. dose of stress agent that led to survival of 50% of the cells, was then calculated using Probit analysis as implemented in NCSS software (NCSS, Kaysville, UT). ED50 values for glucose withdrawal were calculated in a similar manner to estimate the level of glucose or rotenone associated with a 50% reduction in cellular metabolic activity.

Acknowledgements

We wish to thank William Kohler and Melissa Han for development of cell lines and preparation of cell pellets. Supported by NIH AG047745, AG023122, AG047200, DK089503, DK097153, and Life Extension Foundation.

Funding Statement

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

Funding Information

This paper was supported by the following grants:

  • National Institutes of Health DK097153 to Charles F Burant.

  • National Institutes of Health AG047200 to Zhengdong D Zhang, Andrei Seluanov, Vera Gorbunova, Vadim N Gladyshev.

  • National Institutes of Health AG047745 to Vadim N Gladyshev.

  • National Institutes of Health AG023122 to Vadim N Gladyshev.

  • National Institutes of Health DK089503 to Vadim N Gladyshev.

  • Life Extension Foundation to Vera Gorbunova, Andrei Seluanov.

Additional information

Competing interests

The authors declare that no competing interests exist.

Author contributions

SM, Conception and design, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

AU, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

AG, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

Y-MT, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

CFB, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

SR, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

QZ, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

ZDZ, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

AS, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

VG, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

CBC, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

RAM, Conception and design, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

VNG, Conception and design, Analysis and interpretation of data, Drafting or revising the article.

Additional files

Supplementary file 1. Gene expression values.

(A) Raw counts. (B) log10 normalized values.

DOI: http://dx.doi.org/10.7554/eLife.19130.015

elife-19130-supp1.xlsx (4.8MB, xlsx)
DOI: 10.7554/eLife.19130.015
Supplementary file 2. Metabolite levels.

(A) Raw values. (B) log10 normalized values.

DOI: http://dx.doi.org/10.7554/eLife.19130.016

elife-19130-supp2.xlsx (115.9KB, xlsx)
DOI: 10.7554/eLife.19130.016

References

  1. Abegglen LM, Caulin AF, Chan A, Lee K, Robinson R, Campbell MS, Kiso WK, Schmitt DL, Waddell PJ, Bhaskara S, Jensen ST, Maley CC, Schiffman JD. Potential mechanisms for cancer resistance in elephants and comparative cellular response to DNA damage in humans. JAMA. 2015;314:1850–1860. doi: 10.1001/jama.2015.13134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Adelman R, Saul RL, Ames BN. Oxidative damage to DNA: relation to species metabolic rate and life span. PNAS. 1988;85:2706–2708. doi: 10.1073/pnas.85.8.2706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Altschul SF, Madden TL, Schäffer AA, Zhang J, Zhang Z, Miller W, Lipman DJ. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Research. 1997;25:3389–3402. doi: 10.1093/nar/25.17.3389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Austriaco NR, Guarente LP. Changes of telomere length cause reciprocal changes in the lifespan of mother cells in Saccharomyces cerevisiae. PNAS. 1997;94:9768–9772. doi: 10.1073/pnas.94.18.9768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Barja G, Herrero A. Oxidative damage to mitochondrial DNA is inversely related to maximum life span in the heart and brain of mammals. FASEB Journal . 2000;14:312–318. doi: 10.1096/fasebj.14.2.312. [DOI] [PubMed] [Google Scholar]
  6. Brown NP, Leroy C, Sander C. MView: a web-compatible database search or multiple alignment viewer. Bioinformatics. 1998;14:380–381. doi: 10.1093/bioinformatics/14.4.380. [DOI] [PubMed] [Google Scholar]
  7. Busuttil RA, Rubio M, Dollé ME, Campisi J, Vijg J. Oxygen accelerates the accumulation of mutations during the senescence and immortalization of murine cells in culture. Aging Cell. 2003;2:287–294. doi: 10.1046/j.1474-9728.2003.00066.x. [DOI] [PubMed] [Google Scholar]
  8. Camacho C, Coulouris G, Avagyan V, Ma N, Papadopoulos J, Bealer K, Madden TL. BLAST+: architecture and applications. BMC Bioinformatics. 2009;10:421. doi: 10.1186/1471-2105-10-421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Chang HC, Guarente L. Sirt1 mediates central circadian control in the SCN by a mechanism that decays with aging. Cell. 2013;153:1448–1460. doi: 10.1016/j.cell.2013.05.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Chiaromonte F, Miller W, Bouhassira EE. Gene length and proximity to neighbors affect genome-wide expression levels. Genome Research. 2003;13:2602–2608. doi: 10.1101/gr.1169203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Chondrogianni N, Georgila K, Kourtis N, Tavernarakis N, Gonos ES. 20S proteasome activation promotes life span extension and resistance to proteotoxicity in Caenorhabditis elegans. FASEB Journal. 2015;29:611–622. doi: 10.1096/fj.14-252189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Chondrogianni N, Gonos ES. Proteasome activation as a novel antiaging strategy. IUBMB Life. 2008;60:651–655. doi: 10.1002/iub.99. [DOI] [PubMed] [Google Scholar]
  13. Cortopassi GA, Wang E. There is substantial agreement among interspecies estimates of DNA repair activity. Mechanisms of Ageing and Development. 1996;91:211–218. doi: 10.1016/S0047-6374(96)01788-5. [DOI] [PubMed] [Google Scholar]
  14. Coschigano KT, Clemmons D, Bellush LL, Kopchick JJ. Assessment of growth parameters and life span of GHR/BP gene-disrupted mice. Endocrinology. 2000;141:2608–2613. doi: 10.1210/endo.141.7.7586. [DOI] [PubMed] [Google Scholar]
  15. De Guzman JM, Ku G, Fahey R, Youm YH, Kass I, Ingram DK, Dixit VD, Kheterpal I. Chronic caloric restriction partially protects against age-related alteration in serum metabolome. Age. 2013;35:1091–1104. doi: 10.1007/s11357-012-9430-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. de Magalhães JP, Toussaint O. Genage: a genomic and proteomic network map of human ageing. FEBS Letters. 2004;571:243–247. doi: 10.1016/j.febslet.2004.07.006. [DOI] [PubMed] [Google Scholar]
  17. Debrabant B, Soerensen M, Flachsbart F, Dato S, Mengel-From J, Stevnsner T, Bohr VA, Kruse TA, Schreiber S, Nebel A, Christensen K, Tan Q, Christiansen L. Human longevity and variation in DNA damage response and repair: study of the contribution of sub-processes using competitive gene-set analysis. European Journal of Human Genetics : EJHG. 2014;22:1131–1136. doi: 10.1038/ejhg.2013.299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, Batut P, Chaisson M, Gingeras TR. Star: ultrafast universal RNA-seq aligner. Bioinformatics. 2013;29:15–21. doi: 10.1093/bioinformatics/bts635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Edgar RC. Muscle: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Research. 2004;32:1792–1797. doi: 10.1093/nar/gkh340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Eisenberg E, Levanon EY. Human housekeeping genes, revisited. Trends in Genetics : TIG. 2013;29:569–574. doi: 10.1016/j.tig.2013.05.010. [DOI] [PubMed] [Google Scholar]
  21. Essers MA, Weijzen S, de Vries-Smits AM, Saarloos I, de Ruiter ND, Bos JL, Burgering BM. Foxo transcription factor activation by oxidative stress mediated by the small GTPase Ral and JNK. The EMBO Journal. 2004;23:4802–4812. doi: 10.1038/sj.emboj.7600476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Felsenstein J. Phylogenies and the comparative Method. The American Naturalist. 1985;125:1–15. doi: 10.1086/284325. [DOI] [PubMed] [Google Scholar]
  23. Flurkey K, Papaconstantinou J, Harrison DE. The Snell dwarf mutation Pit1(dw) can increase life span in mice. Mechanisms of Ageing and Development. 2002;123:121–130. doi: 10.1016/S0047-6374(01)00339-6. [DOI] [PubMed] [Google Scholar]
  24. Friedman DB, Johnson TE. A mutation in the age-1 gene in Caenorhabditis elegans lengthens life and reduces hermaphrodite fertility. Genetics. 1988;118:75–86. doi: 10.1093/genetics/118.1.75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Fushan AA, Turanov AA, Lee SG, Kim EB, Lobanov AV, Yim SH, Buffenstein R, Lee SR, Chang KT, Rhee H, Kim JS, Yang KS, Gladyshev VN. Gene expression defines natural changes in mammalian lifespan. Aging Cell. 2015;14:352–365. doi: 10.1111/acel.12283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Gentleman R, Falcon S, Sarkar D. Category: Category Analysis. R Package 2013
  27. Ghazi A, Henis-Korenblit S, Kenyon C. Regulation of Caenorhabditis elegans lifespan by a proteasomal E3 ligase complex. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:5947–5952. doi: 10.1073/pnas.0700638104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Gomes AP, Price NL, Ling AJ, Moslehi JJ, Montgomery MK, Rajman L, White JP, Teodoro JS, Wrann CD, Hubbard BP, Mercken EM, Palmeira CM, de Cabo R, Rolo AP, Turner N, Bell EL, Sinclair DA. Declining NAD(+) induces a pseudohypoxic state disrupting nuclear-mitochondrial communication during aging. Cell. 2013;155:1624–1638. doi: 10.1016/j.cell.2013.11.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, Adiconis X, Fan L, Raychowdhury R, Zeng Q, Chen Z, Mauceli E, Hacohen N, Gnirke A, Rhind N, di Palma F, Birren BW, Nusbaum C, Lindblad-Toh K, Friedman N, Regev A. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nature Biotechnology. 2011;29:644–652. doi: 10.1038/nbt.1883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Grube K, Bürkle A. Poly(ADP-ribose) polymerase activity in mononuclear leukocytes of 13 mammalian species correlates with species-specific life span. PNAS. 1992;89:11759–11763. doi: 10.1073/pnas.89.24.11759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Guarente L, Kenyon C. Genetic pathways that regulate ageing in model organisms. Nature. 2000;408:255–262. doi: 10.1038/35041700. [DOI] [PubMed] [Google Scholar]
  32. Ha MK, Soo Cho J, Baik OR, Lee KH, Koo HS, Chung KY. Caenorhabditis elegans as a screening tool for the endothelial cell-derived putative aging-related proteins detected by proteomic analysis. Proteomics. 2006;6:3339–3351. doi: 10.1002/pmic.200500395. [DOI] [PubMed] [Google Scholar]
  33. Hansen M, Chandra A, Mitic LL, Onken B, Driscoll M, Kenyon C. A role for autophagy in the extension of lifespan by dietary restriction in C. Elegans. PLoS Genetics. 2008;4:e24. doi: 10.1371/journal.pgen.0040024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Harper JM, Salmon AB, Leiser SF, Galecki AT, Miller RA. Skin-derived fibroblasts from long-lived species are resistant to some, but not all, lethal stresses and to the mitochondrial inhibitor rotenone. Aging Cell. 2007;6:1–13. doi: 10.1111/j.1474-9726.2006.00255.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Harrison DE, Strong R, Sharp ZD, Nelson JF, Astle CM, Flurkey K, Nadon NL, Wilkinson JE, Frenkel K, Carter CS, Pahor M, Javors MA, Fernandez E, Miller RA. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature. 2009;460:392–395. doi: 10.1038/nature08221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Haussmann MF, Winkler DW, O'Reilly KM, Huntington CE, Nisbet IC, Vleck CM. Telomeres shorten more slowly in long-lived birds and mammals than in short-lived ones. Proceedings. Biological Sciences. 2003;270:1387–1392. doi: 10.1098/rspb.2003.2385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Heilbronn LK, Ravussin E. Calorie restriction and aging: review of the literature and implications for studies in humans. The American Journal of Clinical Nutrition. 2003;78:361–369. doi: 10.1093/ajcn/78.3.361. [DOI] [PubMed] [Google Scholar]
  38. Holzenberger M, Dupont J, Ducos B, Leneuve P, Géloën A, Even PC, Cervera P, Le Bouc Y. Igf-1 receptor regulates lifespan and resistance to oxidative stress in mice. Nature. 2003;421:182–187. doi: 10.1038/nature01298. [DOI] [PubMed] [Google Scholar]
  39. Huang DW, Sherman BT, Lempicki RA. Bioinformatics enrichment tools: paths toward the comprehensive functional analysis of large gene lists. Nucleic Acids Research. 2009a;37:1–13. doi: 10.1093/nar/gkn923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Huang DW, Sherman BT, Lempicki RA. Systematic and integrative analysis of large gene lists using David bioinformatics resources. Nature Protocols. 2009b;4:44–57. doi: 10.1038/nprot.2008.211. [DOI] [PubMed] [Google Scholar]
  41. Jensen LJ, Kuhn M, Stark M, Chaffron S, Creevey C, Muller J, Doerks T, Julien P, Roth A, Simonovic M, Bork P, von Mering C. String 8--a global view on proteins and their functional interactions in 630 organisms. Nucleic Acids Research. 2009;37:D412–416. doi: 10.1093/nar/gkn760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Kamburov A, Wierling C, Lehrach H, Herwig R. Consensuspathdb--a database for integrating human functional interaction networks. Nucleic Acids Research. 2009;37:D623–628. doi: 10.1093/nar/gkn698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Keane M, Semeiks J, Webb AE, Li YI, Quesada V, Craig T, Madsen LB, van Dam S, Brawand D, Marques PI, Michalak P, Kang L, Bhak J, Yim HS, Grishin NV, Nielsen NH, Heide-Jørgensen MP, Oziolor EM, Matson CW, Church GM, Stuart GW, Patton JC, George JC, Suydam R, Larsen K, López-Otín C, O'Connell MJ, Bickham JW, Thomsen B, de Magalhães JP. Insights into the evolution of longevity from the bowhead whale genome. Cell Reports. 2015;10:112–122. doi: 10.1016/j.celrep.2014.12.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Kenyon CJ. The genetics of ageing. Nature. 2010;464:504–512. doi: 10.1038/nature08980. [DOI] [PubMed] [Google Scholar]
  45. Kevei É, Hoppe T. Ubiquitin sets the timer: impacts on aging and longevity. Nature Structural & Molecular Biology. 2014;21:290–292. doi: 10.1038/nsmb.2806. [DOI] [PubMed] [Google Scholar]
  46. Kim EB, Fang X, Fushan AA, Huang Z, Lobanov AV, Han L, Marino SM, Sun X, Turanov AA, Yang P, Yim SH, Zhao X, Kasaikina MV, Stoletzki N, Peng C, Polak P, Xiong Z, Kiezun A, Zhu Y, Chen Y, Kryukov GV, Zhang Q, Peshkin L, Yang L, Bronson RT, Buffenstein R, Wang B, Han C, Li Q, Chen L, Zhao W, Sunyaev SR, Park TJ, Zhang G, Wang J, Gladyshev VN. Genome sequencing reveals insights into physiology and longevity of the naked mole rat. Nature. 2011;479:223–227. doi: 10.1038/nature10533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Kimura M. A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. Journal of Molecular Evolution. 1980;16:111–120. doi: 10.1007/BF01731581. [DOI] [PubMed] [Google Scholar]
  48. Kruegel U, Robison B, Dange T, Kahlert G, Delaney JR, Kotireddy S, Tsuchiya M, Tsuchiyama S, Murakami CJ, Schleit J, Sutphin G, Carr D, Tar K, Dittmar G, Kaeberlein M, Kennedy BK, Schmidt M. Elevated proteasome capacity extends replicative lifespan in Saccharomyces cerevisiae. PLoS Genetics. 2011;7:e1002253. doi: 10.1371/journal.pgen.1002253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Laschober GT, Ruli D, Hofer E, Muck C, Carmona-Gutierrez D, Ring J, Hutter E, Ruckenstuhl C, Micutkova L, Brunauer R, Jamnig A, Trimmel D, Herndler-Brandstetter D, Brunner S, Zenzmaier C, Sampson N, Breitenbach M, Fröhlich KU, Grubeck-Loebenstein B, Berger P, Wieser M, Grillari-Voglauer R, Thallinger GG, Grillari J, Trajanoski Z, Madeo F, Lepperdinger G, Jansen-Dürr P. Identification of evolutionarily conserved genetic regulators of cellular aging. Aging Cell. 2010;9:1084–1097. doi: 10.1111/j.1474-9726.2010.00637.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Lavin SR, Karasov WH, Ives AR, Middleton KM, Garland T. Morphometrics of the avian small intestine compared with that of nonflying mammals: a phylogenetic approach. Physiological and Biochemical Zoology. 2008;81:526–550. doi: 10.1086/590395. [DOI] [PubMed] [Google Scholar]
  51. Liao Y, Smyth GK, Shi W. Featurecounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics. 2014;30:923–930. doi: 10.1093/bioinformatics/btt656. [DOI] [PubMed] [Google Scholar]
  52. Lin SJ, Ford E, Haigis M, Liszt G, Guarente L. Calorie restriction extends yeast life span by lowering the level of Nadh. Genes & Development. 2004;18:12–16. doi: 10.1101/gad.1164804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Löw P. The role of ubiquitin-proteasome system in ageing. General and Comparative Endocrinology. 2011;172:39–43. doi: 10.1016/j.ygcen.2011.02.005. [DOI] [PubMed] [Google Scholar]
  54. Ma S, Lee SG, Kim EB, Park TJ, Seluanov A, Gorbunova V, Buffenstein R, Seravalli J, Gladyshev VN. Organization of the mammalian Ionome according to Organ Origin, Lineage Specialization, and Longevity. Cell Reports. 2015a;13:1319–1326. doi: 10.1016/j.celrep.2015.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Ma S, Yim SH, Lee SG, Kim EB, Lee SR, Chang KT, Buffenstein R, Lewis KN, Park TJ, Miller RA, Clish CB, Gladyshev VN. Organization of the mammalian Metabolome according to organ Function, Lineage Specialization, and Longevity. Cell Metabolism. 2015b;22:332–343. doi: 10.1016/j.cmet.2015.07.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Martin GM. Constitutional genetic markers of aging. Experimental Gerontology. 1988;23:257–270. doi: 10.1016/0531-5565(88)90028-9. [DOI] [PubMed] [Google Scholar]
  57. Martin M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal. 2011;17:10–12. doi: 10.14806/ej.17.1.200. [DOI] [Google Scholar]
  58. McCay CM, Crowell MF, Maynard LA. The Effect of retarded Growth upon the Length of Life Span and upon the Ultimate Body Size. Journal of Nutrition. 1935;10:63–79. [PubMed] [Google Scholar]
  59. McDonnell SR, Hwang SR, Rolland D, Murga-Zamalloa C, Basrur V, Conlon KP, Fermin D, Wolfe T, Raskind A, Ruan C, Jiang JK, Thomas CJ, Hogaboam CM, Burant CF, Elenitoba-Johnson KS, Lim MS. Integrated phosphoproteomic and metabolomic profiling reveals Npm-Alk-mediated phosphorylation of PKM2 and metabolic reprogramming in anaplastic large cell lymphoma. Blood. 2013;122:958–968. doi: 10.1182/blood-2013-01-482026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Meléndez A, Tallóczy Z, Seaman M, Eskelinen EL, Hall DH, Levine B. Autophagy genes are essential for dauer development and life-span extension in C. Elegans. Science. 2003;301:1387–1391. doi: 10.1126/science.1087782. [DOI] [PubMed] [Google Scholar]
  61. Mi H, Muruganujan A, Thomas PD. Panther in 2013: modeling the evolution of gene function, and other gene attributes, in the context of phylogenetic trees. Nucleic Acids Research. 2013;41:D377–386. doi: 10.1093/nar/gks1118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Miller RA, Harrison DE, Astle CM, Fernandez E, Flurkey K, Han M, Javors MA, Li X, Nadon NL, Nelson JF, Pletcher S, Salmon AB, Sharp ZD, Van Roekel S, Winkleman L, Strong R. Rapamycin-mediated lifespan increase in mice is dose and sex dependent and metabolically distinct from dietary restriction. Aging Cell. 2014;13:468–477. doi: 10.1111/acel.12194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Montgomery MK, Buttemer WA, Hulbert AJ. Does the oxidative stress theory of aging explain longevity differences in birds? Ii. Antioxidant systems and oxidative damage. Experimental Gerontology. 2012;47:211–222. doi: 10.1016/j.exger.2011.11.014. [DOI] [PubMed] [Google Scholar]
  64. Moorad JA, Promislow DE, Flesness N, Miller RA. A comparative assessment of univariate longevity measures using zoological animal records. Aging Cell. 2012;11:940–948. doi: 10.1111/j.1474-9726.2012.00861.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Mouchiroud L, Houtkooper RH, Moullan N, Katsyuba E, Ryu D, Cantó C, Mottis A, Jo YS, Viswanathan M, Schoonjans K, Guarente L, Auwerx J. The NAD(+)/Sirtuin pathway modulates longevity through activation of mitochondrial UPR and FOXO signaling. Cell. 2013;154:430–441. doi: 10.1016/j.cell.2013.06.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Murakami S, Salmon A, Miller RA. Multiplex stress resistance in cells from long-lived dwarf mice. FASEB Journal. 2003;17:1565–1566. doi: 10.1096/fj.02-1092fje. [DOI] [PubMed] [Google Scholar]
  67. Pamplona R, Portero-Otín M, Riba D, Requena JR, Thorpe SR, López-Torres M, Barja G. Low fatty acid unsaturation: a mechanism for lowered lipoperoxidative modification of tissue proteins in mammalian species with long life spans. The Journals of Gerontology. Series A, Biological Sciences and Medical Sciences. 2000;55:B286–291. doi: 10.1093/gerona/55.6.B286. [DOI] [PubMed] [Google Scholar]
  68. Pamplona R, Portero-Otín M, Riba D, Ruiz C, Prat J, Bellmunt MJ, Barja G. Mitochondrial membrane peroxidizability index is inversely related to maximum life span in mammals. Journal of Lipid Research. 1998;39:1989–1994. [PubMed] [Google Scholar]
  69. Parrinello S, Samper E, Krtolica A, Goldstein J, Melov S, Campisi J. Oxygen sensitivity severely limits the replicative lifespan of murine fibroblasts. Nature Cell Biology. 2003;5:741–747. doi: 10.1038/ncb1024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Partridge L, Gems D. Mechanisms of ageing: public or private? Nature Reviews. Genetics. 2002;3:165–175. doi: 10.1038/nrg753. [DOI] [PubMed] [Google Scholar]
  71. Perez-Campo R, López-Torres M, Cadenas S, Rojas C, Barja G. The rate of free radical production as a determinant of the rate of aging: evidence from the comparative approach. Journal of Comparative Physiology B: Biochemical, Systemic, and Environmental Physiology. 1998;168:149–158. doi: 10.1007/s003600050131. [DOI] [PubMed] [Google Scholar]
  72. Peters RH. The Ecological Implications Of Body Size. Cambridge University Press; 1986. [Google Scholar]
  73. Pickering AM, Lehr M, Miller RA. Lifespan of mice and primates correlates with immunoproteasome expression. The Journal of Clinical Investigation. 2015;125:2059–2068. doi: 10.1172/JCI80514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Promislow DE. Dna repair and the evolution of longevity: a critical analysis. Journal of Theoretical Biology. 1994;170:291–300. doi: 10.1006/jtbi.1994.1190. [DOI] [PubMed] [Google Scholar]
  75. Revell LJ. phytools: an R package for phylogenetic comparative biology (and other things) Methods in Ecology and Evolution. 2012;3:217–223. doi: 10.1111/j.2041-210X.2011.00169.x. [DOI] [Google Scholar]
  76. Robinson MD, Oshlack A. A scaling normalization method for differential expression analysis of RNA-seq data. Genome Biology. 2010;11:R25. doi: 10.1186/gb-2010-11-3-r25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Röhme D. Evidence for a relationship between longevity of mammalian species and life spans of normal fibroblasts in vitro and erythrocytes in vivo. PNAS. 1981;78:5009–5013. doi: 10.1073/pnas.78.8.5009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Rubinsztein DC, Mariño G, Kroemer G. Autophagy and aging. Cell. 2011;146:682–695. doi: 10.1016/j.cell.2011.07.030. [DOI] [PubMed] [Google Scholar]
  79. Sacher GA. Relation of lifespan to brain weight and body weight in mammals. In: Wolstenholme G. E. W, O'Conner M, editors. Ciba Foundation Symposium - the Lifespan of Animals (Colloquia on Ageing) Chichester: John Wiley & Sons, Ltd; 1959. pp. 115–141. [Google Scholar]
  80. Saitou N, Nei M. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Molecular Biology and Evolution. 1987;4:406–425. doi: 10.1093/oxfordjournals.molbev.a040454. [DOI] [PubMed] [Google Scholar]
  81. Salmon AB, Murakami S, Bartke A, Kopchick J, Yasumura K, Miller RA. Fibroblast cell lines from young adult mice of long-lived mutant strains are resistant to multiple forms of stress. American Journal of Physiology. Endocrinology and Metabolism. 2005;289:E23–29. doi: 10.1152/ajpendo.00575.2004. [DOI] [PubMed] [Google Scholar]
  82. Seim I, Fang X, Xiong Z, Lobanov AV, Huang Z, Ma S, Feng Y, Turanov AA, Zhu Y, Lenz TL, Gerashchenko MV, Fan D, Hee Yim S, Yao X, Jordan D, Xiong Y, Ma Y, Lyapunov AN, Chen G, Kulakova OI, Sun Y, Lee SG, Bronson RT, Moskalev AA, Sunyaev SR, Zhang G, Krogh A, Wang J, Gladyshev VN. Genome analysis reveals insights into physiology and longevity of the Brandt's bat Myotis brandtii. Nature Communications. 2013;4:2212. doi: 10.1038/ncomms3212. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Seim I, Ma S, Zhou X, Gerashchenko MV, Lee SG, Suydam R, George JC, Bickham JW, Gladyshev VN. The transcriptome of the bowhead whale Balaena mysticetus reveals adaptations of the longest-lived mammal. Aging. 2014;6:879–899. doi: 10.18632/aging.100699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Simonsen A, Cumming RC, Brech A, Isakson P, Schubert DR, Finley KD. Promoting basal levels of autophagy in the nervous system enhances longevity and oxidant resistance in adult Drosophila. Autophagy. 2008;4:176–184. doi: 10.4161/auto.5269. [DOI] [PubMed] [Google Scholar]
  85. Slater GS, Birney E. Automated generation of heuristics for biological sequence comparison. BMC Bioinformatics. 2005;6:31. doi: 10.1186/1471-2105-6-31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Smedley D, Haider S, Durinck S, Pandini L, Provero P, Allen J, Arnaiz O, Awedh MH, Baldock R, Barbiera G, Bardou P, Beck T, Blake A, Bonierbale M, Brookes AJ, Bucci G, Buetti I, Burge S, Cabau C, Carlson JW, Chelala C, Chrysostomou C, Cittaro D, Collin O, Cordova R, Cutts RJ, Dassi E, Di Genova A, Djari A, Esposito A, Estrella H, Eyras E, Fernandez-Banet J, Forbes S, Free RC, Fujisawa T, Gadaleta E, Garcia-Manteiga JM, Goodstein D, Gray K, Guerra-Assunção JA, Haggarty B, Han DJ, Han BW, Harris T, Harshbarger J, Hastings RK, Hayes RD, Hoede C, Hu S, Hu ZL, Hutchins L, Kan Z, Kawaji H, Keliet A, Kerhornou A, Kim S, Kinsella R, Klopp C, Kong L, Lawson D, Lazarevic D, Lee JH, Letellier T, Li CY, Lio P, Liu CJ, Luo J, Maass A, Mariette J, Maurel T, Merella S, Mohamed AM, Moreews F, Nabihoudine I, Ndegwa N, Noirot C, Perez-Llamas C, Primig M, Quattrone A, Quesneville H, Rambaldi D, Reecy J, Riba M, Rosanoff S, Saddiq AA, Salas E, Sallou O, Shepherd R, Simon R, Sperling L, Spooner W, Staines DM, Steinbach D, Stone K, Stupka E, Teague JW, Dayem Ullah AZ, Wang J, Ware D, Wong-Erasmus M, Youens-Clark K, Zadissa A, Zhang SJ, Kasprzyk A. The Biomart community portal: an innovative alternative to large, centralized data repositories. Nucleic Acids Research. 2015;43:W589–598. doi: 10.1093/nar/gkv350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Sohal RS, Sohal BH, Orr WC. Mitochondrial superoxide and hydrogen peroxide generation, protein oxidative damage, and longevity in different species of flies. Free Radical Biology and Medicine. 1995;19:499–504. doi: 10.1016/0891-5849(95)00037-X. [DOI] [PubMed] [Google Scholar]
  88. Starke-Reed PE, Oliver CN. Protein oxidation and proteolysis during aging and oxidative stress. Archives of Biochemistry and Biophysics. 1989;275:559–567. doi: 10.1016/0003-9861(89)90402-5. [DOI] [PubMed] [Google Scholar]
  89. Tacutu R, Craig T, Budovsky A, Wuttke D, Lehmann G, Taranukha D, Costa J, Fraifeld VE, de Magalhães JP. Human ageing genomic Resources: integrated databases and tools for the biology and genetics of ageing. Nucleic Acids Research. 2013;41:D1027–1033. doi: 10.1093/nar/gks1155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Tatar M, Kopelman A, Epstein D, Tu MP, Yin CM, Garofalo RS. A mutant Drosophila insulin receptor homolog that extends life-span and impairs neuroendocrine function. Science. 2001;292:107–110. doi: 10.1126/science.1057987. [DOI] [PubMed] [Google Scholar]
  91. Tatusov RL, Koonin EV, Lipman DJ. A genomic perspective on protein families. Science. 1997;278:631–637. doi: 10.1126/science.278.5338.631. [DOI] [PubMed] [Google Scholar]
  92. Tavazoie S, Hughes JD, Campbell MJ, Cho RJ, Church GM. Systematic determination of genetic network architecture. Nature Genetics. 1999;22:281–285. doi: 10.1038/10343. [DOI] [PubMed] [Google Scholar]
  93. Tian X, Azpurua J, Hine C, Vaidya A, Myakishev-Rempel M, Ablaeva J, Mao Z, Nevo E, Gorbunova V, Seluanov A. High-molecular-mass hyaluronan mediates the cancer resistance of the naked mole rat. Nature. 2013;499:346–349. doi: 10.1038/nature12234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Townsend MK, Clish CB, Kraft P, Wu C, Souza AL, Deik AA, Tworoger SS, Wolpin BM. Reproducibility of metabolomic profiles among men and women in 2 large cohort studies. Clinical Chemistry. 2013;59:1657–1667. doi: 10.1373/clinchem.2012.199133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Vernace VA, Arnaud L, Schmidt-Glenewinkel T, Figueiredo-Pereira ME. Aging perturbs 26S proteasome assembly in Drosophila melanogaster. The FASEB Journal . 2007;21:2672–2682. doi: 10.1096/fj.06-6751com. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Weeda G, Donker I, de Wit J, Morreau H, Janssens R, Vissers CJ, Nigg A, van Steeg H, Bootsma D, Hoeijmakers JH. Disruption of mouse ERCC1 results in a novel repair syndrome with growth failure, nuclear abnormalities and senescence. Current Biology. 1997;7:427–439. doi: 10.1016/S0960-9822(06)00190-4. [DOI] [PubMed] [Google Scholar]
  97. Weindruch R, Walford RL, Fligiel S, Guthrie D. The retardation of aging in mice by dietary restriction: longevity, cancer, immunity and lifetime energy intake. The Journal of Nutrition. 1986;116:641–654. doi: 10.1093/jn/116.4.641. [DOI] [PubMed] [Google Scholar]
  98. Western D. Size, life history and ecology in mammals. African Journal of Ecology. 1979;17:185–204. doi: 10.1111/j.1365-2028.1979.tb00256.x. [DOI] [Google Scholar]
  99. Willcox BJ, Donlon TA, He Q, Chen R, Grove JS, Yano K, Masaki KH, Willcox DC, Rodriguez B, Curb JD. FOXO3A genotype is strongly associated with human longevity. PNAS. 2008;105:13987–13992. doi: 10.1073/pnas.0801030105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Wishart DS, Jewison T, Guo AC, Wilson M, Knox C, Liu Y, Djoumbou Y, Mandal R, Aziat F, Dong E, Bouatra S, Sinelnikov I, Arndt D, Xia J, Liu P, Yallou F, Bjorndahl T, Perez-Pineiro R, Eisner R, Allen F, Neveu V, Greiner R, Scalbert A. HMDB 3.0--The human Metabolome Database in 2013. Nucleic Acids Research. 2013;41:D801–807. doi: 10.1093/nar/gks1065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Wood JG, Rogina B, Lavu S, Howitz K, Helfand SL, Tatar M, Sinclair D. Sirtuin activators mimic caloric restriction and delay ageing in metazoans. Nature. 2004;430:686–689. doi: 10.1038/nature02789. [DOI] [PubMed] [Google Scholar]
  102. Zhang Q, Nogales-Cadenas R, Lin J-R, Zhang W, Cai Y, Vijg J, Zhang ZD. Systems-level analysis of human aging genes shed new light on mechanisms of aging. Human Molecular Genetics. 2016:ddw145. doi: 10.1093/hmg/ddw145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Zhang Y, Shao Z, Zhai Z, Shen C, Powell-Coffman JA. The HIF-1 hypoxia-inducible factor modulates lifespan in C. Elegans. PloS One. 2009;4:e6348. doi: 10.1371/journal.pone.0006348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Zhu J, Adli M, Zou JY, Verstappen G, Coyne M, Zhang X, Durham T, Miri M, Deshpande V, De Jager PL, Bennett DA, Houmard JA, Muoio DM, Onder TT, Camahort R, Cowan CA, Meissner A, Epstein CB, Shoresh N, Bernstein BE. Genome-wide chromatin state transitions associated with developmental and environmental cues. Cell. 2013;152:642–654. doi: 10.1016/j.cell.2012.12.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
eLife. 2016 Nov 22;5:e19130. doi: 10.7554/eLife.19130.018

Decision letter

Editor: Amy J Wagers1

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

Thank you for submitting your article "Cell culture-based profiling across mammals reveals DNA repair and metabolism as determinants of species longevity" for consideration by eLife. Your article has been favorably evaluated by Janet Rossant as the Senior Editor and three reviewers, one of whom is a member of our Board of Reviewing Editors. The following individuals involved in review of your submission have agreed to reveal their identity: Daniel Promislow (Reviewer #2) and Yousin Suh (Reviewer #3).

The reviewers have discussed the reviews with one another and the Reviewing Editor has drafted this decision to help you prepare a revised submission.

Summary:

The manuscript by Gladyshev and colleagues seeks to correlate gene expression and metabolite signatures in cultured fibroblasts from 16 different mammals with longevity phenotypes of the respective mammals, including maximal lifespan, body size, female time to maturity, and stress resistance. Using RNAseq and mass spec, the authors show that the constructed gene expression profiles reflect phylogenetic relationships, and identify subsets of genes and metabolites that correlate, positively or negatively, with multiple longevity traits (but not necessarily body size). The study presents a massive amount of work, particularly in compiling the various datasets and developing the informatics pipelines to specify species-specific ortholog sets and assess the robustness of the results, which will certainly be of interest to the community of researchers studying the biology of aging. All three reviewers agreed that the study provides interesting insights into aging and longevity, and should be of broad interest. However, as detailed below, a number of concerns were also raised, relating in large part to insufficient clarity in the current version with respect to the authors' methods and the limitations of their approach and data.

Essential revisions:

1) As this fibroblast culture system is the cornerstone of the authors' approach, it is essential that they provide clear and complete details regarding the experimental strategies used to isolate and culture these cells. In particular, detailed answers to the following key questions must be available in the present manuscript:

A) What age and sex were the animals from which the fibroblasts were isolated (e.g., was% max lifespan of donor animals matched for the different species? Were they sex-matched?)?

B) From what region of the body were fibroblasts isolated? This is important since other studies have shown that fibroblasts retain a regional gene expression program (e.g., hox patterning) after isolation.

C) What is the media system used, and strategy for cell passaging? How were these culture conditions chosen, and how can it be determined if they are optimal for all species (e.g., the authors recently published on a specific sensitivity to oxygen tension for fibroblasts from lab mice, but not other rodent species). This information should be very clearly articulated and discussed in the manuscript text itself, since it forms the basis for the experimental system used.

D) At what passage number were gene expression and metabolite profiles generated?

E) What is the in vitro proliferation rate of the different fibroblast isolates, and does this correlate with any of the gene expression profiles or longevity traits analyzed?

2) It should also be discussed that in vitro cell culture conditions dramatically change chromatin architecture (Zhu et al., Cell 2013), and so gene expression patterns. The gene expression signature detected may simply reflect the ability of cells to differentially adjust to the in vitro conditions, and may contribute to the multiplex stress resistance. This caveat should be discussed in the text.

3) It is not clear whether the common 9,389 gene orthologs that were reliably detected across the 15 species are comparable to the number of expressed genes detected by RNA-seq in fibroblasts in the 5 species with annotated genomes. The base line information on how many transcripts are detected in the 5 species should be available as they may provide a way to estimate the number of potentially missing orthologs, due to sequence divergence, some of which may be critical contributors to longevity through drastic functional alterations of gene products. The text should be clarified to address this point.

4) It is not clear what the baseline gene set was for the enrichment analysis of the 827 top hits (Figure 3D). Was it 9,389 genes or whole gene sets? And if the latter, which species? The data should be presented to show the enriched pathways in the top hits using the whole gene sets vs. 9,389 genes as baseline. This is because if DNA repair pathways are already enriched among the 9,389 genes, the conclusion that it is a longevity-associated molecular signature can be misleading. It is formerly possible that genes in these ancient pathways (such as DNA repair or glucose metabolism) may be more conserved sequence-wise and therefore more enriched among the orthologs, and thus among top hits. The text and data presentation should be clarified to address this point.

5) Subsection “Longevity trait variation across mammal”. The authors need to provide more information on how they calculated residuals from body mass. The reference to Ma et al. 2015b gives allometric equations, but no information is provided regarding where those equations come from. The description in Ma et al. (residual LS = LS/(a x Massb)) implies that residuals were taken from a least-squares regression of log(LS) vs. log(Mass) with an intercept of log(a) and a slope of b. But this is not mentioned in the present manuscript, and must be inferred from the Ma et al. manuscript. If a least-squares regression approach was used, one important assumption is that the residuals are normally distributed. The data for Figure 1 (MLres and FTMres) do not look normally distributed. If they are not, a general linear model with an appropriate link function must be applied.

6) The authors use maximum lifespan as a metric of aging, and thus should add to the text of their manuscript a discussion of the concerns that have been raised regarding this statistic (e.g. Moorad et al. 2012 Aging Cell). In particular, it does not provide a measure of aging per se, is quite sensitive to small sample size, and is not in itself under direct selection.

7) Analyses of primate and bird species (subsection “Validation of amino acid patterns in primate and bird fibroblasts”) does not appear to be size corrected. Numerous studies have shown strong size effects of life span not only in primates, but also in birds. Therefore, size effects should be addressed here as well.

8) Subsection “Evaluation of amino acid levels in bird and primate fibroblast cell lines”. For the bird/primate analysis, the authors state, "when two or more […] features were annotated as corresponding to the same amino acid, we tabulated the degree of association from the feature most strongly correlated with lifespan among the species studied." This approach will necessarily bias the result in favor of suggesting a stronger relationship between lifespan and metabolite levels than might be true, and this caveat must be mentioned.

eLife. 2016 Nov 22;5:e19130. doi: 10.7554/eLife.19130.019

Author response


Essential revisions:

1) As this fibroblast culture system is the cornerstone of the authors' approach, it is essential that they provide clear and complete details regarding the experimental strategies used to isolate and culture these cells. In particular, detailed answers to the following key questions must be available in the present manuscript:

We agree that these are very important points. We have updated the Methods section to incorporate these reviewers’ comments and suggestions. Additional comments are provided below.

A) What age and sex were the animals from which the fibroblasts were isolated (e.g., was% max lifespan of donor animals matched for the different species? Were they sex-matched?)?

Fibroblasts were isolated from post-pubertal adults. Although ages of the wild-captured donors could not be reliably determined, it is a good assumption that most were young, i.e. their age was substantially less than the maximal lifespan that could be achieved by protected or captured members of the species. For mice, for example, maximum lifespan in the laboratory is approximately 3.5 to 4 years, but mice in their natural environment typically live about six months. Thus, most wild-captured adults were likely to be well below half of the species-specific maximal recorded longevity. The sex of the donors was not recorded. Variations in donor age or sex are unlikely to have generated false positive findings in our design. If such variations (age or sex, for example) did have any effects on our endpoints, such effects would have increased Type II error (false negatives), but not have led us to false positive claims (Type I error), unless these demographic variables were confounded with species life history variables. For example, if most of the animals captured from short-lived species were males, and most of those captured from longer-lived species were females, this could in principle create a false positive error by confounding species lifespan with gender effects, but such a scenario seems unlikely.

B) From what region of the body were fibroblasts isolated? This is important since other studies have shown that fibroblasts retain a regional gene expression program (e.g., hox patterning) after isolation.

We used an identical procedure to prepare cells. Fibroblasts were taken from sun-protected abdominal skin, as previously reported (Harper et al., Aging Cell 2007). They were therefore consistent across all our species.

C) What is the media system used, and strategy for cell passaging? How were these culture conditions chosen, and how can it be determined if they are optimal for all species (e.g., the authors recently published on a specific sensitivity to oxygen tension for fibroblasts from lab mice, but not other rodent species). This information should be very clearly articulated and discussed in the manuscript text itself, since it forms the basis for the experimental system used.

The detailed culture conditions are now incorporated under “Cell culture” in the Materials and methods section. In particular, these cultures were kept under low-oxygen conditions (3% O2) after thawing to minimize selection for resistance to O2 toxicity. All of the preparations used for RNA and metabolite testing were similar in degree of expansion, all were at the earliest possible passage number from the original skin biopsy, and were expanded under low O2 conditions.

The cell lines were all exposed to the same culture conditions. We do not know if the conditions were "optimal" for rapid growth of cells from each species, but clearly the use of identical culture conditions is needed to avoid confounding effects due to species origin with possible effects of culture variations. We recently showed that cells from wild mice and other wild rodents are much less sensitive to O2 toxicity than cells of laboratory mice (Patrick et al., Aging 2016), but in any case we tried to minimize any such effects by use of 3% O2 and by doing all analyses on cells at early passage numbers.

D) At what passage number were gene expression and metabolite profiles generated?

See answer to point C) above and the revised Materials and methods section. Cryopreserved cells were prepared at passage 2 (counting the first transfer of confluent cells as passage 1); thawed aliquots of 106 cells were then allowed to expand to 30 x 106 cells, an increase of approximately 4 or 5 doublings.

E) What is the in vitro proliferation rate of the different fibroblast isolates, and does this correlate with any of the gene expression profiles or longevity traits analyzed?

We do not know the "in vitro proliferation" rate of our cell lines. This would vary in complex ways with seeding density, glucose and serum levels, degree of confluence at subculture, and days after seeding at which the proliferation was to be measured (lag phase, log expansion phase, or near-confluence, for example). It is possible that cell lines from different species might differ in some of these parameters, and that these differences could contribute to species differences in gene expression in a pattern that reflects longevity of the species tested, but evaluating this idea would be a complex undertaking. Nevertheless, the fibroblasts used for gene expression were collected at similar degree of confluency (with ~ 30 x 106cells), so any variations in growth rate will be less likely to influence the gene expression pattern.

2) It should also be discussed that in vitro cell culture conditions dramatically change chromatin architecture (Zhu et al., Cell 2013), and so gene expression patterns. The gene expression signature detected may simply reflect the ability of cells to differentially adjust to the in vitro conditions, and may contribute to the multiplex stress resistance. This caveat should be discussed in the text.

We agree with the reviewers and have included this point in the Discussion. We used cells from the second/third passage for the RNA-seq and metabolite measurements in order to minimize in vitro culture artifacts, but it is still possible that some of the features reflect the differential responses to in vitro culture stress.

3) It is not clear whether the common 9,389 gene orthologs that were reliably detected across the 15 species are comparable to the number of expressed genes detected by RNA-seq in fibroblasts in the 5 species with annotated genomes. The base line information on how many transcripts are detected in the 5 species should be available as they may provide a way to estimate the number of potentially missing orthologs, due to sequence divergence, some of which may be critical contributors to longevity through drastic functional alterations of gene products. The text should be clarified to address this point.

We appreciate these comments and have provided the additional information in the main text and in Figure 1—source data 1C “Mapping to Genomes”.

Figure 1—source data 1C compares the results of mapping to the publicly available genomes and to our ortholog sets. 5 species (big brown bat, little brown bat, guinea pig, chinchilla, and deer mouse) have publicly available genomes in ENSEMBL and NCBI. For these species, we mapped the reads to the full genomes and the associated annotations, and counted the number of expressed genes (>10 counts).

The number of annotated genes in these species ranged between ~14,000 to ~16,000. Such variations are likely due to the different pipelines used for gene annotation. Among these annotated genes, ~10,000-11,000 had >10 counts (the same criteria for the expressed orthologs). This is similar to the 9389 orthologs we identified, indicating that our list of expressed orthologs is likely to have captured at least 80% of the expressed genes.

4) It is not clear what the baseline gene set was for the enrichment analysis of the 827 top hits (Figure 3D). Was it 9,389 genes or whole gene sets? And if the latter, which species? The data should be presented to show the enriched pathways in the top hits using the whole gene sets vs. 9,389 genes as baseline. This is because if DNA repair pathways are already enriched among the 9,389 genes, the conclusion that it is a longevity-associated molecular signature can be misleading. It is formerly possible that genes in these ancient pathways (such as DNA repair or glucose metabolism) may be more conserved sequence-wise and therefore more enriched among the orthologs, and thus among top hits. The text and data presentation should be clarified to address this point.

We have clarified the pathway enrichment procedure in the Materials and methods section.

The background for pathway enrichment was based on the 16,816 unique protein-coding genes identified as Mouse Reference (i.e. Step 1 in Figure 2A, after removing highly repetitive and highly similar sequences). The pathways were based on Mus musculus. We have now also included pathway enrichment statistics performed using only the 9389 expressed orthologs as background (Table 1—source data 1G,H). In both cases, the top enriched pathways remained statistically significant.

Regarding the comment “It is formerly possible that genes in these ancient pathways (such as DNA repair or glucose metabolism) may be more conserved sequence-wise and therefore more enriched among the orthologs, and thus among top hits”, we would like to discuss it in terms of the following aspects:

1) We agree with the reviewers that the genes in the DNA repair and glucose metabolism pathways are likely well conserved across the species. In fact, our additional analysis suggested that many of our ML-associated genes are “housekeeping genes” (Table 1—source data 1I). This is consistent with the notion that lifespan regulation across the species likely requires the co-ordination of a number of basic biological and cellular pathways. Since the longevity variation falls along a continuum across the different mammalian species, it is logical that the underlying evolution forces act on a set of pathways that are common (and conserved) across these species (i.e. the “public” pathways).

2) To the extent that our results are biased by sequence conservation (i.e. that we identified the DNA repair and glucose metabolism pathways only because of their high sequence conservation, and that we miss out many other important genes due to our ortholog identification strategy), we do not believe this is the case for the following reasons:

A) First, for those species with publicly available genomes, when we mapped the reads to the public genomes and used the public annotations, only ~10,000 to ~11,000 genes were expressed (based on a criterion of > 10 counts; same criterion in our ortholog filtering step), even though ~14,000 to ~16,000 genes were annotated (Figure 1—source data 1C). This is mainly because the expression of many genes are tissue-specific, so they are simply not expressed in fibroblasts. In comparison, we identified 9389 expressed ortholog sets across all 15 species. Our 9389 orthologs should therefore represent at least 80% of all the expressed genes in fibroblasts of each species, and the correlation was between 0.93 to 0.98 for the read counts.

B) To address the issue of sequence conservation and divergence, we calculated the median DNA distance (using Kimura 2-parameters distance) for each ortholog set (Figure 2—figure supplement 1C and D). In particular, among our 9389 expressed orthologs, there was no statistical difference in DNA distance between those orthologs that showed correlation with longevity and those orthologs that did not (Wilcoxon Rank Sum Test p value = 0.32; see Author response image 1). In other words, the degree of sequence conservation or divergence is not linked to the degree of correlation with longevity. There are many other genes that are either more conserved or less conserved in sequence than the genes in DNA repair or glucose metabolism; but these other genes did not show up in the top hits because of their lack of expression correlation to longevity, not because of their sequence conservation or divergence per se.

Author response image 1.

Author response image 1.

DOI: http://dx.doi.org/10.7554/eLife.19130.017

5) Subsection “Longevity trait variation across mammal”. The authors need to provide more information on how they calculated residuals from body mass. The reference to Ma et al. 2015b gives allometric equations, but no information is provided regarding where those equations come from. The description in Ma et al. (residual LS = LS/(a x Massb)) implies that residuals were taken from a least-squares regression of log(LS) vs. log(Mass) with an intercept of log(a) and a slope of b. But this is not mentioned in the present manuscript, and must be inferred from the Ma et al. manuscript.

We have now included an additional paragraph (“Life history data of the species”) under the Methods section to provide more information on the allometric equation. The ML equation, MLres = ML/(4.88×AW0.153), was based directly on the documentation of the AnAge database (http://genomics.senescence.info/help.html#anage). The relevant paragraph was quoted here:

“For mammals, also included is the maximum longevity (tmax) residual, expressed as a percentage of the expected maximum longevity calculated from the adult body size (M) and derived from the mammalian allometric equation: tmax = M0.153. This is useful to identify species that live longer than expected for their body size. Cetaceans were excluded because we have less confidence in their longevity records, obtained from studies in the wild often using indirect methods, than in those from other mammalian taxa.”

The FTM equation, FTMres = FTM/(78.1×AW0.217), was calculated by us, by linear regression using the FTM and body mass records of 1330 mammalian species in the AnAge database.

If we had used the same procedure to calculate the MLres equation, we would obtain MLres = ML/(4.36×AW0.158), which is comparable to the AnAge documentation.

If a least-squares regression approach was used, one important assumption is that the residuals are normally distributed. The data for Figure 1 (MLres and FTMres) do not look normally distributed. If they are not, a general linear model with an appropriate link function must be applied.

We had tested the normalcy of the data using Shapiro test. The results were as follows (if p-value < 0.05, then the data is not normally distributed):

Trait (on log10 scale)

Shapiro test p-value

ML

0.8401

FTM

0.0753

MLres

0.7969

FTMres

0.1258

Therefore, none of the longevity traits violated normalcy assumption.

Furthermore, we also tested the normalcy of the gene expression data and metabolite data. Using Shapiro test, 89% of the genes and 90% of the metabolites did not violate normalcy assumption on log10 scale.

6) The authors use maximum lifespan as a metric of aging, and thus should add to the text of their manuscript a discussion of the concerns that have been raised regarding this statistic (e.g. Moorad et al. 2012 Aging Cell). In particular, it does not provide a measure of aging per se, is quite sensitive to small sample size, and is not in itself under direct selection.

Thank you. We have included this point in the Discussion section. Additional parameters such as mean adult lifespan and 90th percentile of longevity (Moorad et al., 2012) should be used in place of ML in the long run, although at the moment such records are available for only a limited number of species.

7) Analyses of primate and bird species (subsection “Validation of amino acid patterns in primate and bird fibroblasts”) does not appear to be size corrected. Numerous studies have shown strong size effects of life span not only in primates, but also in birds. Therefore, size effects should be addressed here as well.

We have now expanded Table 2 to include the body-mass adjusted lifespan correlation. We would like to point out that:

A) Removal of body mass weakened the observed correlation (Table 2). This is likely due to the inherent strong correlation between maximum lifespan and body mass. Indeed, the same weakening is also observed among the rodent fibroblasts (Table 2—source data 1), indicating the two datasets are in fact consistent.

B) The same weakening of correlation is also observed in our previous studies (Ma et al., 2015 Cell Metabolism and Ma et al., 2015 Cell Reports). This is in fact expected, given the strong correlation between maximum lifespan and body mass. It is highly debatable whether one could completely disentangle these 2 traits, since they are likely under the same natural selection force.

C) Whether the body-mass should be removed in cross-species longevity studies is an area where experts in comparative biology hold widely divergent views. Some (e.g. John Speakman) think that "uncorrected" data are not interpretable; another set, equally experienced (e.g. Tony Hulbert), thinks that an attempt to "correct" for mass will create a major risk for missing key associations, since mass and lifespan are so strongly confounded across species. Others suggest that one should report both unadjusted and adjusted values so that readers can see whichever set of results they feel is most valid.

D) Such associations, whether or not corrected for mass, are only a first step towards more definite, causal, hypotheses. Some remain significant under both ML and MLres, while others do not. Nevertheless, they can all be potential candidates for further experimental verification.

8) Subsection “Evaluation of amino acid levels in bird and primate fibroblast cell lines”. For the bird/primate analysis, the authors state, "when two or more […] features were annotated as corresponding to the same amino acid, we tabulated the degree of association from the feature most strongly correlated with lifespan among the species studied." This approach will necessarily bias the result in favor of suggesting a stronger relationship between lifespan and metabolite levels than might be true, and this caveat must be mentioned.

This selection reflects, in part, uncertainties in the annotation of specific features in the untargeted metabolomics protocol used for the bird and primate species. In most cases where two or more of the annotated features were thought to represent the same amino acid, there was very good agreement. The table below gives the key information on this point. It lists each amino acid for which there was more than one annotated feature, and gives the p-values for the relationship with species lifespan for each such feature.

Amino acid

p-values

Arginine

0.03, 0.1, 0.2

Glutamic acid

0.01, 0.2

Tryptophan

0.001, 0.001, 0.01

Methionine

0.03, 0.1

Phenylalanine

0.001, 0.003, 0.01

Proline

0.001, 0.001

Valine

0.001, 0.001

In four cases (tryptophan, phenylalanine, proline, valine), all such features show a significant association with lifespan. For the other three, the features that do not meet the traditional p = 0.05 criterion have two-sided p-values between 0.05 and 0.2, with the same sign of the slope coefficient.

We agree with the reviewer that our decision to present the feature with the best fit to the MLS regression could in principle create bias and thus increase Type I error, but in this case there are several reasons for confidence: the close fit between each of the nominal "replicate" features within the bird/primate data set, the good agreement across each of the 10 amino acids evaluated, and the excellent agreement to the relationship seen for the rodent cell lines evaluated with an entirely different method.

Associated Data

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

    Supplementary Materials

    Figure 1—source data 1. Species and samples used in the current study.

    (A) Species and traits information. Life history traits of adult weight (AW, in grams), maximum lifespan (ML, in years), and female time to maturity (FTM, in days) of these species were obtained from Anage database (Tacutu et al., 2013). Since the life history data were not available for meadow vole, the data of a related species Microtus arvalis were used instead. Maximum lifespan residual (MLres) and female time to maturity residuals (FTMres) were computed based on the allometric equations MLres = ML/(4.88×AW0.153) and FTMres = FTM/(78.1×AW0.217), respectively. (B) RNA sequencing and read mapping to ortholog sets. Read mapping statistics are based on STAR. De novo assembly was performed by Trinity. (C) Read mapping to publically available genomes. For the species with publicly available genomes, the reads were also aligned to the full genomes for mapping rate comparison. The numbers of annotated genes were based on the published annotations. (D) Metabolite profiling. Metabolite profiling was performed on selected species only.

    DOI: http://dx.doi.org/10.7554/eLife.19130.003

    DOI: 10.7554/eLife.19130.003
    Table 1—source data 1. Phylogenetic regression of gene expression against longevity traits.

    Regression against (A) Adult Weight; (B) Maximum Lifespan; (C) Female Time to Maturity; (D) Maximum Lifespan Residual; and (E) Female Time to Maturity Residual. ‘coef.all’, ‘p value.all’, and ‘q value.all’ refer to the regression slope, p value, and FDR-adjusted q value using all the species. ‘p value.robust’ and ‘q value.robust’ refer to the statistics after removing the potential outlier point. ‘p value.max’ and ‘q value.max’ refer to the maximal (least significant) regression p value and q value when each one of the species was left out, one at a time. Only genes with p value.robust<0.01 and p value.max<0.05 are shown. (F) Top hits identified by two or more longevity traits. The p value.robust against each of the four longevity traits (ML, FTM, MLres, and FTMres) as well as adult weight (AW) are shown. These genes were the input for pathway enrichment analysis. Pathway enrichment analysis of genes showing (G) positive and (H) negative correlation with longevity traits. Enrichment was performed using DAVID with default settings. Only the top 10 clusters are shown. (I) System level analyses of gene functions. The numbers of shared genes between longevity associated genes (either positively or negatively or both) and human aging genes, essential genes, transcription factor genes, and housekeeping genes are shown. The enrichment p value was calculated by Fisher’s exact test with two different background gene sets.

    DOI: http://dx.doi.org/10.7554/eLife.19130.009

    DOI: 10.7554/eLife.19130.009
    Table 2—source data 1. Phylogenetic regression of metabolite levels against longevity traits.

    Regression against (A) Adult Weight; (B) Maximum Lifespan; (C) Female Time to Maturity; (D) Maximum Lifespan Residual; and (E) Female Time to Maturity Residual. ‘coef.all’, ‘p value.all’, and ‘q value.all’ refer to the regression slope, p value, and FDR-adjusted q value using all the species. ‘p value.robust’ and ‘q value.robust’ refer to the statistics after removing the potential outlier point. ‘p value.max’ and ‘q value.max’ refer to the maximal (least significant) regression p value and q value when each one of the species was left out, one at a time. Only genes with p value.robust < 0.01 and p value.max < 0.05 are shown. (F) Top hits identified by two or more longevity traits. The p value.robust against each of the four longevity traits (ML, FTM, MLres, and FTMres) as well as adult weight (AW) are shown. These metabolites were the input for pathway enrichment analysis. Pathway enrichment analysis of metabolites showing (G) positive and (H) negative correlation with longevity traits. Enrichment was performed based on hypergeometric statistics. (I) Top hits identified by two or more longevity traits, using cut-off of p value.robust < 0.05. The p value.robust against each of the four longevity traits (ML, FTM, MLres, and FTMres) as well as adult weight (AW) are shown.

    DOI: http://dx.doi.org/10.7554/eLife.19130.014

    DOI: 10.7554/eLife.19130.014
    Supplementary file 1. Gene expression values.

    (A) Raw counts. (B) log10 normalized values.

    DOI: http://dx.doi.org/10.7554/eLife.19130.015

    elife-19130-supp1.xlsx (4.8MB, xlsx)
    DOI: 10.7554/eLife.19130.015
    Supplementary file 2. Metabolite levels.

    (A) Raw values. (B) log10 normalized values.

    DOI: http://dx.doi.org/10.7554/eLife.19130.016

    elife-19130-supp2.xlsx (115.9KB, xlsx)
    DOI: 10.7554/eLife.19130.016

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

    RESOURCES