Introduction

Aging is a certainty in our largely uncertain lives. It is a process in which multiple organs and tissues gradually lose physiological integrity, followed by functional impairment and eventually death of the individual1. The molecular mechanisms underlying aging are not fully understood, despite the enormous amount of findings and theories that have emerged in the past decades. The current hypotheses encompass genetic predisposition, calorie restriction, mitochondrial dysfunction, telomere attrition, genomic instability and many others2,3,4,5,6. As there is also no unanimous agreement on fundamental issues such as whether aging is genetically programmed7,8, the ultimate cause of aging and the interconnections among various aging mechanisms remain to be established. On the contrary is the fact that aging is a major risk factor for many complex diseases such as cardiovascular disease, cancer, Type 2 diabetes, Alzheimer’s disease and Parkinson’s disease9,10,11,12,13. Given the rapidly expanding aging population world-wide14, aging research is increasingly important as it holds the promise for unravelling the secrets of longevity and for bringing new solutions to the treatment of age-related diseases.

With the advent of various high throughput technologies, it is now feasible to measure an individual’s panomics (including transcriptome, metabolome, epigenome, etc.) at a reasonable cost15. The rich information in panomic data brings enormous opportunities to the aging research field. For example, using methylation data, Horvath et al. defined a molecular clock composed of 353 CpG sites that could accurately predict the human age16. By examining the transcriptome changes in the aging neocortex and cerebellum in mice, Lee et al. observed genes associated with inflammatory responses, oxidative stress and reduced neurotrophic support in both brain regions17. The AGEMAP project which profiled gene expression in 16 tissues in mice also identified age-associated genes and revealed tissue specific aging patterns18. By comparing the transcriptional profiles in mice to those of other species (human, flies and worms), genes involved in the electron transport chain showed common age regulation in all four species. A large number of human tissue age-gene expression association studies have been performed in various tissues (e.g., brain, muscle, blood and kidney)19,20,21,22,23,24,25. However, the previous gene expression based studies only examined a rather limited number of tissue types. Due to difference in sample collections, platforms used for profiling and data processing procedures, it is difficult to compare and combine the findings from these studies. The GTEx project provides RNA-Seq based transcriptome profiles in more than 40 tissues from hundreds of human donors of various ages, making it one of the largest single data sets with the most comprehensive tissue types for studying the genetics of human tissue gene expression and age-associated gene expression26. Particularly, since multiple tissues are collected from the same individuals, cross-tissue analysis of age-associated gene expression changes becomes feasible. For simplicity we subsequently refer to such genes as “aging genes”.

In this work, we first identify aging gene signatures in nine GTEx tissues and explore their functional characteristics. We then study the synchronization of age-related gene expression changes across different tissues. We also study the connections between tissue aging and complex diseases. In addition, we examine the aging gene signatures across different species and compare our results with other related aging studies.

Results

Identification of tissue specific age-associated genes in the GTEx data

The GTEx project (v3, accessed in December 2012) provided 1,641 whole transcriptome profiles in more than 40 tissues from nearly two hundred post-mortem human donors26. Nine tissues had sample sizes of greater than 80, namely, the subcutaneous adipose, tibial artery, left ventricle heart, lung, skeletal muscle, tibial nerve, skin, thyroid and whole blood. We considered these nine tissues in our study and omitted other tissue types which had fewer samples. The sample age and gender distributions were plotted in Fig. 1, in which the age distributions of all samples, female samples and male samples were shown in the left, middle and right columns, respectively and each row corresponds to a tissue type. Overall, donors’ chronological ages ranged from 20 to 70. A customized regression model combined with bootstrapping was used to define age-gene expression associations (see Methods).

Figure 1
figure 1

Age distribution of donors in nine tissues.

The histograms of donor age distribution for all samples, females only and males only.

Briefly, we followed the GTEx consortium’s practice of pre-processing the gene expression data with a slight modification26. We corrected gender, the top three genotype principal components (PCs) to reduce the impact of population structure and a few top gene expression PCs that did not significantly correlate with age (p-value > 0.05). Correcting gene expression PCs allows us to remove potential confounding factors such as the batch effect27. Multiple alternative models of correcting expression PCs were considered and compared (see details in Methods). Since low expressed genes are usually more vulnerable to measurement errors28, we removed 20% low expressed aging genes. We bootstrapped the samples for 100 times and compiled our final aging gene list by including genes whose expression levels were significantly associated with age in at least 50 runs. The readers are referred to Methods for the details.

We summarized the numbers of age-associated genes in all nine tissues in Table 1 and the detailed information for all the 41,298 genes was provided in Supplementary Data S1. As shown in Table 1, the number of age-associated genes in the nine tissues ranged from 3 to 3,287. The largest number (n = 3,287) was observed in whole blood; while only 12 and 3 age-associated genes were identified in skin and thyroid, respectively. The large difference in the number of age-associated genes observed in the GTEx data is consistent with previous observations in mouse18 and human22. To estimate the number of false positives that could be contained in our aging genes, we permuted sample ages for 1,000 times and repeated the aging gene identification procedure on the permutated data (see Methods). The results were summarized in the column named “Permutation” in Table 1. As shown in Table 1, the frequencies of identifying equal or more age-associated genes in the permuted datasets are very small (no greater than 5 times except for skin and thyroid) and the numbers of false positives are small compared to the number of findings in the real data in most tissues (skin and thyroid are the two exceptions), indicating that majority of the identified gene-age associations in the seven tissues are not due to random chance.

Table 1 Number of age-associated genes in 9 human tissues and permutation test.

Because of the small number of age-associated genes and high false positive rates in the skin and thyroid tissues, we excluded them and only performed further analysis on seven tissues. To visualize the expression pattern of the inferred aging genes, a heatmap was generated for each tissue (Supplementary Fig. S1). As shown in Fig. 2a, for the adipose tissue, the samples were clearly clustered into two groups based on the Euclidian distance with “Ward” measurement29. The left-side group contained many younger individuals compared to the right-side group. The apparent separation of “young” and “old” group was observed in all seven tissues and it is of note that the groupings were different in different tissues (Supplementary Fig. S1). A student’s t-test on the age groups, “young” and “old”, generated significant results (p-values less than 3.59 × 10−3) in all seven tissues (Supplementary Table S1). The age distributions of “young” and “old” samples in the seven tissues were provided in Supplementary Fig. S2.

Figure 2
figure 2

Age-associated gene expression in subcutaneous adipose tissue.

(a) Heatmap of 1,134 age-associated genes (row) on 94 samples (column). Colors represent normalized gene expression values with blue for low expression and red for high expression. The age of each individual is displayed at the bottom and also illustrated in color bar at the top with dark green for young and yellow for old. (b) Scatter plot of 2 representative age-associated gene expression patterns PYH1N1 and EIF5AL1 in adipose tissue. Pearson-R value in the title represents the Pearson correlation coefficient between gene expression and age across all samples. The solid blue triangles plot male samples and solid red circle female samples. Similarly, the blue and red lines denote the regression lines for male and female samples, respectively.

In addition to the grouping on the age axis, the aging genes were also grouped into up-regulated and down-regulated genes, which correspond to positive and negative signs of the coefficient of “age” term in the regression model. In four out of the seven tissues, we saw more up-regulated aging genes than down-regulated ones (heart, lung and blood are exceptions) (Table 1). To visualize the age-gene expression association at single gene level, we selected two genes with either strong positive or negative age-association (PYH1N1 and EIF5AL1) in adipose tissue and showed the scatter plots in Fig. 2b. The Pearson correlation coefficients of these two genes are 0.63 (p-value = 7.43 × 10−12) and −0.57 (p-value = 2.89 × 10−9), respectively. As can be seen in Fig. 2b, the age-gene expression association is evident for these two genes and there is no apparent difference between male (blue) and female (red). The scatter plots of the top 100 age-associated genes in all seven tissues are provided in Supplementary Data S2.

Functional annotation of aging genes points to a large collection of biological processes

To obtain a functional overview, we annotated the up- and down-regulated aging genes separately, using David tools30. Due to space limitation, we showed a subset of the top representative annotations in Table 2 and provided a complete list in Supplementary Data S3. The aging gene signatures are significantly enriched for a wide spectrum of Gene Ontology (GO) terms and pathways. The most frequently appeared category is mitochondrion. In five out of seven tissues (adipose, artery, heart, lung and blood), the enrichment was all seen in the down-regulated aging genes. This result supports the central role of mitochondria in human aging. Mitochondria dysfunction in aging has been observed in multiple model organisms and is among the most recognized aging theories1,31,32,33. For example, a systemic RNA interference (RNAi) screen for gene inactivation that increases lifespan in worms showed a 10-fold overrepresentation of genes encoding mitochondrial proteins34. Many other functions known to associate with aging were also observed, e.g., down-regulation of electron transport chain (in adipose and heart) and up-regulation of cell death and inflammation response (both in artery). It is of note that several disease pathways, e.g., “hsa05016: Huntington’s disease”, “hsa05012: Parkinson’s disease” and “hsa05010: Alzheimer’s disease” are significantly enriched in aging gene signatures in multiple tissues and these neurodegenerative diseases are also known to be age-dependent12,13,35.

Table 2 Function enrichment of up- and down-regulated age-associated genes in 7 tissues.

Besides the well-known age-related functions, we also saw multiple processes that are less known for their involvement in aging. For example, we observed the up-regulation of cell adhesion in nerve and up-regulation of ion binding in adipose. All these significantly over-represented biological functions suggest that human aging is an extremely broad and complex process, with both common and specific, up- and down-regulated collections of biological activities in various tissues.

Synchronization of multi-tissue age-related gene expression changes in human

As shown in Fig. 2a and Supplementary Fig. S1, although the chronological ages of the tissue donors are largely consistent with the “young” or “old” clustering, there are some apparent “outliers”. For example, the 15th adipose sample (from the left) in Fig. 2a is from an individual with an old chronological age of 70, but is clustered with the “young” group where most donors range in age from 20 to 50. This indicated that the molecular age of adipose tissue from this individual is about 30 years younger than its chronological age. Sample mislabelling can be excluded as a possible cause for such “outliers”, due to the accurate sample matching using variants called from RNA-Seq data and DNA genotype data. In fact, this finding is consistent with previous observations in the mouse data from the AGEMAP project18. Although AGEMAP concluded that different tissues in the same mouse tend to have coordinated aging, they also observed “outliers” or mouse tissues that clustered into different aging patterns. They hypothesized that individual mice could be composed of a mosaic of tissues with different physiological ages.

One question we asked is whether age-related expression changes in different tissues are coordinated. It is of note that our study design is not longitudinal, therefore we are unable to study the coordinated aging gene expression changes among tissues in single individuals. Instead, we studied this at the population level with the assumption that common patterns of age-related expression changes exist in human population and can be observed in our samples. To answer the aforementioned question, we estimated tissues’ apparent ages or their relative ages (ranks) in the population and then we calculated their covariation in the population for each tissue-pair. Specifically, we considered two methods: (1) an unsupervised method based on principal component analysis (PCA) and (2) a supervised learning method using the Elastic Net regression36.

PCA analysis

Briefly, we calculated the PCs of aging gene expression in each tissue. In all seven tissues, the first PC (PC1) captured the dominant variance of the aging gene expression and sample coordinate on PC1 was highly correlated with the sample’s chronological age with a mean Spearman correlation coefficient of 0.55 (Supplementary Fig. S3). Therefore we only considered PC1 to estimate the relative rank of tissue’s apparent age.

As we performed such calculation in each tissue separately, different tissues from the same individual may have different age ranks among donors. We then calculated the Spearman correlation coefficient of two tissues’ apparent ages across all the individuals, named as “co-aging coefficient”. A high co-aging coefficient indicates that the tissue pair has tight synchronized age-related expression changes in the population, so that if we randomly select an individual and profile the tissue pair, when one tissue appears to be young, the other tissue has a high chance of being young too, or vice versa. We plotted all pair-wise tissue correlations in Fig. 3a. As shown in Fig. 3a, high correlations were seen among lung, blood and heart (e.g., the correlation between lung and blood was 0.73), whereas much less correlations were observed among other tissues (e.g., the correlation between muscle and blood was only 0.27) (Fig. 3b,c).

Figure 3
figure 3

Aging synchronization in multiple tissues.

(a) Heatmap of tissue co-aging in adipose, artery, heat, lung, muscle, nerve and blood tissues. The number in each square in the lower triangle indicates the co-aging coefficient between a tissue pair; and the sample number for each tissue is also presented in the top of the sample column. (b) Scatter plot of age rank correlation for two highly correlated tissues lung and blood and the regression line. Spearman-R value represents the Spearman correlation coefficient between the ranks defined by the two tissues across all samples. (c) Scatter plot of age rank correlation for two relatively uncorrelated tissues muscle and blood and the regression line. (d) Scatter plot of age rank correlation for heart, lung and blood; the axis ranges from 0 to 60 indicating the rank of each sample (in total 59 donors with data in all these 3 tissues); the chronological ages are marked for some representative samples and the outliers are highlighted in red. (e) Correlation between age rank deviation and age in artery and nerve. The shaded area indicates the confidence interval of the regression line. “p-value” indicates the p-value for regression coefficient being deviated from 0. (f) Correlation between rank deviation and age in lung and blood.

One potential cause for observing the synchronization among tissue specific age-related gene expression changes is cross-tissue contamination, in which case gene expression profiled from one tissue is a mixture of gene expression from multiple tissues (e.g., heart tissues might be contaminated with blood). To evaluate the level of contamination, we performed a clustering of samples based on gene expression using (1) all genes and (2) all aging genes. We observed that the samples collected from the same tissue were grouped tightly together in either case (see Supplementary Fig. S4a and S4b). This indicated that cross-tissue contamination may not be the main cause for the strong correlation of age-related gene expression changes amongst certain tissues.

The tight co-aging amongst the lung, heart and blood tissues is better visualized in Fig. 3d, in which we mapped each overlapping sample (across the three tissues) into a ball in 3 dimensional space according to its rank (based on the coordinate of PC1 projection) in each tissue. It is clear that most samples showed synchronized aging in these three tissues (ball points in grey color), as they are aligned with the diagonal line connecting (0, 0, 0) and (60, 60, 60) and positioned closely to their chronological ages. A couple of outliers are also seen, as they are far off the diagonal line indicating large deviation from donors’ chronological ages. We labelled samples with deviation p-values less than 0.05 as outliers and colored them in red (see Methods for details). We also observed that deviation has a positive correlation with chronological age (see Supplementary Table S2), as we saw greater age deviations in old individuals than those in young individuals. This is particularly true for tissue pairs with relatively weak synchronization. For example, for the artery and nerve tissues (Fig. 3e), the deviation of sample age rank increases significantly as age increases (p-value = 2.04 × 10−4); but for the lung and blood tissues (Fig. 3f), although the correlation remains positive, it is not significant (p-value = 0.18).

Predicted age using Elastic Net

As an alternative approach to PCA, we predicted tissue age using Elastic Net (EN) and performed tissue co-aging analysis using the predicted ages. Briefly, we randomly divided samples into 10 subgroups of equal size and predicted sample age in each subgroup using data from the other 9 subgroups. To reduce bias due to random sampling, we repeated the process for 100 times and the mean of the predicted age in these runs was used as sample’s predicted age (see Supplementary Methods for details). The tissue co-aging patterns based on EN were plotted in Supplementary Fig. S5. As can be seen, the pattern is different from the one obtained by PCA method. Using EN, the most correlated tissue pair is nerve and artery (Spearman correlation coefficient of 0.76), while in the PCA analysis, apparent ages of heart, lung and blood are highly correlated. As a supervised learning method, Elastic Net is designed to minimize the difference between the predicted age and chronological age in the training samples. As a consequence, we observed that the Spearman correlation between Elastic Net predicted age and chronological age was consistently higher than that in the PCA results for all tissues (Supplementary Table S3). The apparent “outliers” seen in the PCA analysis were predicted to be less dramatic by the EN. For example, the 70 years old adipose tissue was ranked at 13th position in all the 94 samples based on its PC1 projection (smaller rank corresponds to younger age), while based on EN, it was ranked more to the donor’s chronological age (36th position). In Supplementary Fig. S5, we also listed the rooted mean square error (RMSE) of EN prediction for each tissue. With current small sample size, EN showed large errors, e.g., artery and nerve showed RMSEs of 6.75 and 7.20 respectively, while blood tissue showed the largest RMSE of 11.94.

Despite the difference in co-aging patterns observed in PCA and EN analyses, both methods showed that the aging of artery and nerve is strongly correlated with chronological age and this correlation is much smaller for blood, suggesting different tissues may have different levels of deviation from the individual’s chronological age (Supplementary Table S3).

Tissue specific link between aging genes and complex disease genes

Aging is a known risk factor for many diseases, such as cardiovascular disease, cancer, arthritis, Alzheimer’s disease, Parkinson’s disease and Type 2 diabetes37,38,39. The functional enrichment in the previous section indicated that aging gene signatures have clear connections with some age-related diseases like Alzheimer’s disease. To comprehensively evaluate the tissue specific connections between aging and diseases, we compiled a large disease gene list containing 234 disease/trait categories by merging two datasets, the NIH GWAS40 and OMIM41 catalog (see Methods for details).

Using this gene set, we determined the enrichment of the disease genes in tissue specific aging gene signatures considering the up- and down-regulated gene signatures separately. A large number of diseases showed tissue specific connections with aging genes. To visualize the result, we selected the top 10 diseases/traits that showed the most significant enrichment for aging genes in each tissue. By considering the unique diseases/traits in all seven tissues, we obtained 55 diseases that showed significant over-representation in up-regulated aging genes and 53 diseases that showed significant over-representation in down-regulated aging genes. The results for up-regulated aging genes and down-regulated aging genes were shown in Fig. 4a,b, respectively and the full enrichment analysis results were provided in Supplementary Data S4. As shown in Fig. 4, in some cases, the over-representation of disease genes in tissue specific aging gene signatures was observed in the tissue types that are commonly considered as the disease affected tissue. For example, the down-regulated aging gene signature in the lung shows most significant enrichment in genes associated with chronic obstructive pulmonary disease (COPD)-related biomarkers and pulmonary function, with p-values of 2.51 × 10−5 and 3.40 × 10−5, respectively. The total cholesterol associated genes and obesity-related traits are significantly over-represented in the adipose tissue down-regulated aging genes with p-values of 8.32 × 10−3 and 2.84 × 10−2, respectively. It is of note that some of these p-values may not survive the multiple testing correction and could represent false positive results. In addition, a large number of immune-mediated inflammatory disease genes are over-represented in the blood and artery aging gene signatures, including Crohn’s disease, inflammatory bowel disease, multiple sclerosis, rheumatoid arthritis and ulcerative colitis. Chronic low-grade systemic inflammation is a common manifestation of aging42,43 and our results further support the strong connection between aging and inflammatory diseases in the human population. On the other hand, some enrichment results are less intuitive, for example, disease genes of age-related macular degeneration (AMD), a disease of the eye, appeared to be over-represented in the up-regulated aging genes of the lung (Fig. 4a). Overall our results suggest that the connections between aging and diseases are very complex. Although we observed some direct connections between disease and tissue type, many connections could be indirect and thus undetectable from simple enrichment analysis. Aging and complex diseases could also be fundamentally different, as some individuals can have a long period of disease free aging life.

Figure 4
figure 4

Correlation between age and common diseases in multiple tissues.

(a) Overlap enrichment between up-regulated aging genes and disease genes in GTEx tissues and (b) Overlap enrichment between down-regulated aging genes and disease genes in GTEx tissues. The color depth indicates the normalized negative logarithm of the p-value of the Fisher’s exact test for overlapping between disease and aging genes in specific tissue.

Aging genes in mouse are different from human aging genes from GTEx

Mouse models have been widely used to study human diseases. Although multiple studies have employed various mouse models to study aging18,44, it is not clear whether at the molecular level, mouse aging is comparable to human aging. To address this question, we compared the human aging genes derived from the GTEx data with mouse aging genes obtained from the AGEMAP project18. The AGEMAP project identified mouse age-associated genes in nine tissues: adrenals, cerebellum, eye, gonads, heart, lung, spleen, spinal cord and thymus, among which two tissues (i.e., heart and lung) were also profiled by GTEx.

To do a cross species comparison, we first obtained the homologous mapping from homologene database in NCBI (released in 12/14/2012) ( ftp://ftp.ncbi.nih.gov/pub/HomoloGene/). Based on this mapping, 6,454 and 6,576 homologous (human/mouse) gene pairs were identified in the heart and lung data, respectively from both GTEx and AGEMAP (Supplementary Data S5). Among the 6,454 homologous genes, there are 346 aging genes (p-value < 0.001, similar to18) in the human heart (Supplementary Data S5) and 18 aging genes in the mouse heart18. Similarly, there are 324 aging genes in the human lung (Supplementary Data S5) and 66 aging genes in the mouse lung18. There is only one common age-associated gene DAZAP1 in both human and mouse heart tissue (p-value = 0.63, one-tail Fisher’s exact test). DAZAP1 (DAZ Associated Protein 1) is required for normal growth and spermatogenesis in mice45,46 and is deleted in many azoospermic and severely ligospermic men47. A recent study showed that DAZAP1 regulates the splicing of Crem, Crisp2 and Pot1a transcripts48. Similarly, there are only three common age-associated genes PTPN6, CDKN1C and CLIP1 (mapped to Hcph, Cdkn1c and Rsn in mouse) in lung tissue, the p-value for the one-tail Fisher’s exact test is 0.64.

Overall, our results suggest a very large difference in aging genes between the human and mouse, a finding consistent with results from previous studies18,49. The large difference in the aging gene pattern has also been observed in other animals like chimpanzee50, suggesting the aging process is indeed less well conserved across species as compared to some other biological processes51. It is of note that the study designs including sample size, gender distribution, age distribution and experiment conditions are quite different between the AGEMAP and GTEx studies. This might contribute in part to the observed large difference and further investigation is required to ensure that the dissimilarity of the two species is not due to design artefacts.

Comparison with other human age-focused gene expression studies

Glass et al. performed an analysis on the Multiple Tissue Human Expression Resource (MuTHER) data to identify aging genes in skin, blood, brain and adipose tissues22. Three tissues are related to GTEx tissues, namely, adipose (subcutaneous fat), skin and lymphoblastoid cell lines (LCLs) from blood.

We applied our model to the MuTHER dataset and compared the identified aging genes with those from GTEx. As shown in Supplementary Table S4, 157 common aging genes were identified in adipose tissue (enrichment p-value = 1.51 × 10−4). In contrast, only 4 common aging genes were identified in skin and blood tissues, respectively (enrichment p-values were both 0.17). The common aging genes were listed in Supplementary Data S6.

We also compared aging genes we derived from the MuTHER dataset using our method with those reported by GLASS et al.22. They showed a significant overlap with p-values of 1.39 × 10−177, ≈0 (< 4.9 × 10−324) and 2.06 × 10−7 respectively in adipose, skin and blood (see Supplementary Data S7). This implies that the large difference of aging genes in the skin and blood between MuTHER and GTEx data is inherent to the data themselves, rather than driven by the difference in the analytic methods. It is of note that in MuTHER project, peripheral blood samples were collected and LCLs were generated through EBV-mediated transformation of the B-lymphocyte component, while in GTEx, the whole blood tissues were used for gene expression profiling. Therefore, it is likely that the difference in aging genes from LCLs vs GTEx whole blood is at least partially due to the sample difference. In addition, the skin tissues were collected from different parts of the body in GTEx (sun exposed lower leg) and MuTHER (relatively photo-protected infra-umbilical skin). It indicates that even the same tissue may have different aging patterns given a different environment such as sun light exposure. Other factors, such as the different gene expression platforms may also contribute to the gross differences in aging genes observed in the different human studies.

Discussion

In the present study, we present a holistic view of human aging in multiple tissues by analysing the GTEx data. From GO and pathway enrichment analyses, mitochondrion biology is highlighted as the most commonly regulated biological process associated with age. We also observed many other biological processes associated with age such as DNA repair, electron transport chain and apoptosis. With extensive involvement of a large number of biological processes during aging, an important task is to identify the connections amongst these aging components and reveal the causal relationship among them and identify the key driver genes as candidates for anti-aging drug development.

Through tissue co-aging analysis, we show that tissue aging as reflected by the age-related gene expression changes is potentially synchronized at different levels. In PCA analysis, vital tissues like lung, heart and whole blood showed tighter co-aging pattern compared to other tissues like muscle. The different levels of tissue age-related expression change synchronization may reflect the different selection pressures on the functional connections among tissues in the early developmental stage which extends into the late stage of our lifespan.

The high correlation between blood and heart/lung provides a hint that circulation system could play a role in synchronizing tissue functions with respect to aging. Recent studies have shown that blood from young mice and possibly GDF11, a circulating transforming growth factor–β (TGF-β) family member can reverse cardiac hypertrophy52, stimulate brain vascular remodelling and increase neurogenesis in aging mice53.

It is of note that the co-aging pattern derived from PCA and Elastic Net are different. This difference may be caused by the limited sample size that is currently available for EN prediction. A more robust evaluation is possible when GTEx project completes with more than 900 individuals being profiled.

Our study identifies some apparent connections between aging and complex diseases. However, there are a large number of connections between tissue specific aging and diseases that are much less obvious and some are hard to interpret. Clearly, a more detailed study is required to fully understand the mechanistic details of all the connections between aging and diseases.

Methods

GTEx data

GTEx data (v3, December 2012 release) provides expression levels of 41,298 genes in nine human tissues: subcutaneous adipose, tibial artery, left ventricle heart, lung, skeletal muscle, tibial nerve, skin (from sun exposed lower leg), thyroid and whole blood. The sample size of each tissue ranges from 83 to 156 (see Table 1). Since GTEx consortium paper26 provides full information on sample collection, RNA collection, RNA-Seq experiment, gene expression estimation, quality control and gene expression normalization, we did not reproduce such information here.

Linear regression model for aging gene detection

In each tissue, we modelled gene expression using the following linear regression model:

where, Yij is the expression level of gene j in sample i, Agei denotes the age of sample i, Sexi denotes the sex of sample i, Genotypeik(1 ≤ k ≤ 3) denotes the value of the k-th principal component value of genotype profile for the i-th sample, PCik(1 ≤ k ≤ N) denotes the value of the k-th principal component value of gene expression profile for the i-th sample, N is the total number of top PCs under consideration, εij is the error term, βj is the regression intercept (for gene j), γj is the age regression coefficient, δj is the sex regression coefficient, μjk(1 ≤ k ≤ 3) is the regression coefficient for the k-th genotype PC and αjk(1 ≤ k ≤ N) is the regression coefficient for the k-th gene expression PC. For each gene j, a least square approach was used to estimate the regression coefficients. If γj was significantly deviated from 0, gene j was considered to be age-associated. Gene j was up-regulated with age if γj > 0 and down-regulated if γj < 0. We performed the false discovery rate (FDR) adjustment on the p-values using Benjamini Hochberg method54 and an FDR less than 0.05 was used as the significance threshold throughout the paper unless otherwise specified.

In addition, we also removed 20% low expressed aging genes. Specifically, for each aging gene, we ranked samples based their expression levels in descending order and calculated the mean expression levels in the top 25% samples. We then ranked all the aging genes according to this mean expression value in descending order and removed the bottom 20% aging genes.

Correcting for confounding factors based on principal component analysis

Correcting confounding factors is usually indispensable in revealing the true relationship between gene expression change and aging. Besides sex, major principal components (PCs) of genotype and gene expression profile in sample space are also frequently used as confounding factors in gene expression analysis to boost true signal detection. For example, removal of the top PCs has led to a significant increase in the number of expression quantitative trait loci (eQTLs) identified55,56. To remove potential confounding factors such as the batch effect, we adopted an approach similar to Pickrell et al.27. It is of note that we only considered top 3 genotype PCs (similar to GTEx eQTL study26) and top 5 gene expression PCs throughout this study.

Since age could be one of the top gene expression PCs or correlate with them, directly correcting these PCs is not suitable for our study. Thus, we tested nine models (namely, M1–M9) to correct possible confounding factors: (1) M1: no correction of any confounding factors; (2) M2: correcting gender and the top 3 genotype PCs; (3) M3-M7: correcting gender, top 3 genotype PCs and the gene expression PCs correlated with age with the Pearson correlation coefficient less than a threshold of 0.1, 0.2, 0.3, 0.4 and 0.5, respectively; (4) M8: correcting gender, the top 3 genotypes and the combination of gene expression PCs that delivered the largest number of age-associated genes; and (5) M9: correcting gender, the top 3 genotype PCs and gene expression PCs that do not significantly (p-value > 0.05) correlate with age. The Pearson correlation between the top 5 gene expression PCs and age was listed in Supplementary Table S5 and the number of aging gene identified after removing 20% low expressed genes was listed in Supplementary Table S6. It can be seen that the chronological age was significantly correlated with top PCs in a few tissues (e.g., PC1 in adipose, artery, heart, lung and blood) (Supplementary Table S5). In addition, confounding factors correction had some influence on the number of inferred aging genes especially for heart, lung and blood (Supplementary Table S6).

To help us evaluate these models, we performed an enrichment analysis with GenAge genes57 using 20,059 GTEx protein coding genes as background (Supplementary Data S8). GenAge provides a manually curated list of 298 genes that presumably regulate the human aging process (accessed on Feb 20th, 2015). The enrichment analysis results of the nine models were shown in Supplementary Table S7, in which we also demonstrated the effect of removing 20% low expressed age-associated genes. In all models, age-associated genes were more significantly overlapped with GenAge genes after removing low expressed genes, suggesting that this filtering may help to refine the true age-associated genes. The model generating the most significant p-value is M9, by which we inferred a total of 7,925 unique protein-coding age-associated genes from all nine GTEx tissues. The number of overlap genes between these aging genes and GenAge genes is 173 with a p-value of 2.69 × 10−11 (Supplementary Table S7).

Bootstrapping, permutation analysis and the effect of sample size

To further ensure our age-associated genes are not sensitive to a particular input sample set, we bootstrapped the samples (with replacement) for 100 times. In each run, we identified age-associated genes from the bootstrapped samples using M9. A gene is in the final age-associated gene list if: (1) it is an age-associated gene in more than 50 bootstrap runs and (2) it is an age-associated gene using the full sample set.

In addition, we performed permutation analysis to estimate the fraction of false positives in our findings. Specifically, we randomly permuted the ages of samples and repeated the test using M8 for 1,000 times. We counted the number of tests in which more age-associated genes were identified and removed the tissues whose number was larger than 5. It is of note that M8 delivers greater number of aging genes than M9, therefore provides an upper bound value of FDR.

To estimate the impact of sample size, we randomly selected samples of sizes 10 to 150 with 10 additional samples added each time for each tissue and repeated this process 100 times using M8 (see Supplementary Fig. S6). As expected, large sample sizes increases the power of identifying age-associated genes (e.g., more than 30 fold increase in the number of detected age-associated genes in whole blood when sample size increases from 10 to 150).

Synchronization pattern of multi-tissue aging in humans

To study the co-aging of multiple tissues, a PCA method was performed to reduce the dimensionality of the gene expression data. We compared different tissues based on the first PC of gene expression. The co-aging coefficient of two tissues was defined as the Spearman correlation coefficient between the coordinates of samples on the first PC (for the two tissues) (see Fig. 3a).

To estimate the aging deviation of a sample in n tissues, we mapped each sample into an n dimensional Euclidean space with each coordinate being its rank (based on its coordinate on PC1) among the overlapping samples in a tissue. Similarly, we also ranked the samples by their chronological age. Let (r1, r2,···, rn) be the rank vector of a sample on n tissues and r be its rank on chronological age. We defined rank deviation d of a sample as .

Visualizing tissue co-aging in 3D space

We choose heart, lung and blood as an example for visualizing tissue co-aging in 3D space. After rank deviation was calculated for each sample, we estimated the corresponding p-value assuming that the distance has a normal distribution and considered samples with p-values less than 0.05 as outliers. Jmol ( http://www.jmol.org/) was used for the visualization (see Fig. 3d).

Assembly of disease gene list and disease-aging gene link detection

The disease genes were retrieved from two sources: NIH Genome-Wide Association Studies (GWAS) Catalog40 (accessed on Aug 13, 2014) and OMIM (Online Mendelian Inheritance in Man, accessed on Aug 13, 2014)41. We only considered genes in the GWAS Catalog with p-value < 5 × 10−8, a generally accepted threshold for genome-wide significance. Clustering and manual curation were used to merge genes in GWAS and OMIM (See Supplementary Data S9 and S10). We only considered diseases with at least 5 genes. We then performed a Fisher’s exact test between the disease genes and aging genes in each tissue. Aging genes with FDR ≤ 0.02 were used for testing age-disease overlap enrichment (p-values less than 0.05 were considered significant). In addition, we separated the up- and down-regulated genes with age. To visualize the result, we selected the top 10 most significant diseases in each tissue, which resulted in 55 unique diseases for up-regulated or 53 for down-regulated aging genes. The normalized for each disease-tissue pair was plotted in Fig. 4.

Data access

The GTEx genotype and gene expression data were downloaded from dbGap under dbGaP Study Accession number phs000424.v3.p1. MuTHER gene expression data was obtained from http://www.ebi.ac.uk/arrayexpress/experiments/E-TABM-1140/, accessed on 2-26-2015.

Additional Information

How to cite this article: Yang, J. et al. Synchronized age-related gene expression changes across multiple tissues in human and the link to complex diseases. Sci. Rep. 5, 15145; doi: 10.1038/srep15145 (2015).