Skip to main content
eLife logoLink to eLife
. 2024 Jul 22;13:e94007. doi: 10.7554/eLife.94007

Multi-omic analysis of bat versus human fibroblasts reveals altered central metabolism

N Suhas Jagannathan 1,2,, Javier Yu Peng Koh 1,, Younghwan Lee 1, Radoslaw Mikolaj Sobota 3, Aaron T Irving 4,5, Lin-fa Wang 6, Yoko Itahana 1,, Koji Itahana 1,, Lisa Tucker-Kellogg 1,2,
Editors: Pankaj Kapahi7, Pankaj Kapahi8
PMCID: PMC11262796  PMID: 39037770

Abstract

Bats have unique characteristics compared to other mammals, including increased longevity and higher resistance to cancer and infectious disease. While previous studies have analyzed the metabolic requirements for flight, it is still unclear how bat metabolism supports these unique features, and no study has integrated metabolomics, transcriptomics, and proteomics to characterize bat metabolism. In this work, we performed a multi-omics data analysis using a computational model of metabolic fluxes to identify fundamental differences in central metabolism between primary lung fibroblast cell lines from the black flying fox fruit bat (Pteropus alecto) and human. Bat cells showed higher expression levels of Complex I components of electron transport chain (ETC), but, remarkably, a lower rate of oxygen consumption. Computational modeling interpreted these results as indicating that Complex II activity may be low or reversed, similar to an ischemic state. An ischemic-like state of bats was also supported by decreased levels of central metabolites and increased ratios of succinate to fumarate in bat cells. Ischemic states tend to produce reactive oxygen species (ROS), which would be incompatible with the longevity of bats. However, bat cells had higher antioxidant reservoirs (higher total glutathione and higher ratio of NADPH to NADP) despite higher mitochondrial ROS levels. In addition, bat cells were more resistant to glucose deprivation and had increased resistance to ferroptosis, one of the characteristics of which is oxidative stress. Thus, our studies revealed distinct differences in the ETC regulation and metabolic stress responses between human and bat cells.

Research organism: Other

Introduction

Bats display many characteristics that set them apart from other mammals, including the capacity for wing-powered flight, low rates of cancer incidence (Seluanov et al., 2018), high longevity quotient (Austad and Fischer, 1991; Austad, 2010; Wilkinson and South, 2002), and ability to carry many viruses (as a reservoir) without ill health (Wang et al., 2011). Each of these traits has distinct metabolic requirements and can affect overall metabolic activity. For example, flight is an energy-intensive process that requires high metabolic rates and ATP production (Thomas and Suthers, 1972; Maina, 2000). Resistance to cancer depends on multiple factors such as reactive oxygen species (ROS) management, DNA repair (Huang et al., 2016; Huang et al., 2019), and the ability to efflux genotoxic compounds (Koh et al., 2019), all of which have metabolic underpinnings. Longevity and disease resistance both require tolerance to metabolic and oxidative stresses, and the ability to dampen inflammasome activation (Ahn et al., 2019; Kacprzyk et al., 2017).

Multiple studies have documented individual aspects of metabolic regulation in bats. Metabolic rates of bats during flight (ATP production) have been documented to be approximately three times greater than basal metabolic rates in other mammals of similar size (Thomas and Suthers, 1972). In most other mammalian species, increased ATP production also creates increased production of ROS in the mitochondria, which can eventually lead to DNA and cellular damage (Buffenstein et al., 2008), and activate inflammasome responses. On the other hand, low amounts of ROS are a part of homeostasis and have been linked to beneficial effects on survival in multiple contexts (Mittler, 2017; Di Meo et al., 2016). It is conceivable that the homeostatic amount of ROS for a healthy cell is species-specific, as different species may have different ways of coping with the adverse effects of ROS accumulation, e.g., ROS-induced DNA damage and lipid peroxidation, or may activate downstream pathways such as inflammasomes at different levels of ROS. Hence, for bats to be able to have a high metabolic rate without concomitant cellular/DNA damage would require one of the following to be true – improved decoupling of ATP production from ROS production (reduced leakage of electrons, so less ROS is produced), improved antioxidant defense to neutralize generated ROS, or improved repair of damage resulting in fewer deleterious effects of ROS (e.g. less inflammasome activation).

Different studies have proposed different mechanisms of ROS tolerance or antioxidant defense in bats including lower hydrogen peroxide production (Brunet-Rossinni, 2004; Brunet-Rossinni and Austad, 2004; Podlutsky et al., 2005; Ungvari et al., 2008), improved DNA repair (Foley et al., 2018), higher expression of heat shock proteins (Chionh et al., 2019) and/or a drug efflux factor, ABCB1 (Koh et al., 2019), and positive selection for efficient mitochondria. While the reasons could be multifactorial, no studies have performed a systems-wide comparison of mitochondrial metabolism between bats and higher mammals such as humans. Such characterization of the basal energy metabolism of bats and how it is different from other mammals such as humans could shed light on mitochondrial activity/ROS management and hint toward metabolic factors that underlie/support desirable traits in bats, e.g., longevity, low cancer/mutation rates, and disease tolerance. Toward this goal, we compared the basal metabolism of cells from black flying fox fruit bat Pteropus alecto (P. alecto) and human cells. P. alecto is a member of the Pteropodidae family and is among the largest fructivore bats in the world. It has a lifespan of over 20 years and is documented to co-exist with lethal zoonotic viruses like Hendra and Nipah viruses. Using a primary lung fibroblast cell line that our group had established earlier from P. alecto (PaLung), we conducted a comparison between PaLung cells and human primary fibroblasts WI-38 cells to elucidate fundamental metabolic differences in the mitochondria.

Since bats and humans are very different species, it is likely that data from any one high-throughput platform (e.g. transcriptomics) would show many differences. Hence, to identify consistent differences in the metabolic regulation of P. alecto and humans, we looked for concordance between multiple high-throughput platforms: whole-cell transcriptomics, mitochondrial proteomics, and whole-cell metabolomics. To integrate the different omics results, we use constraint-based flux sampling, which is a computational modeling technique that simulates metabolic flux patterns using existing knowledge about metabolic network connectivity and topology. Constraint-based flux modeling has been used previously for comparing metabolic phenotypes across cancers (Aurich et al., 2017), understanding metabolic regulation of macrophage polarization (Bordbar et al., 2012), studying ischemia-reperfusion injury (Chouchani et al., 2014), characterizing microbiomes (Jansma and El Aidy, 2021; Ezzamouri et al., 2023), optimizing metabolite production (Patil et al., 2004), and understanding metabolic contributors of disease pathology, e.g., diabetes (Ravi and Gunawan, 2021).

Here, our results show that PaLung cells have differences in basal metabolism that resemble ischemia, including the possibility of low or reverse activity of Complex II in the electron transport chain (ETC). Finally, we characterized the response of bat cells to cellular stresses such as oxidative stress, nutrient deprivation, and a type of cell death related to ischemia, viz. ferroptosis, and results were consistent with our prediction of ischemic-like basal metabolism in PaLung cells.

Results

Transcriptomics identifies differences in oxidative phosphorylation between PaLung cells and WI-38 cells

To understand the differences in cellular-scale metabolism between PaLung cells (P. alecto) and WI-38 cells (Homo sapiens), we performed whole-cell transcriptomics on the two cell lines. Transcriptomics detected a total of 21,952 mRNA transcripts in bat PaLung cells and 58,830 transcripts in human WI-38 cells, respectively. Since the bat genome is not as well annotated as the human genome, we performed downstream differential expression (DE) analysis using the set of 14,986 common transcripts found in both PaLung and WI-38 cells. DE analysis was performed using the DEseq pipeline, which was modified to normalize for the different transcript lengths in both species (see Materials and methods, Figure 1A, Figure 1—figure supplement 1A). This method yielded a total of 6247 transcripts that passed our cutoff thresholds (|log fold change| ≥ 1 and FDR < 0.05, Figure 1B, Figure 1—figure supplement 1B). Because there is no standard way of normalizing RNAseq data for inter-species comparison, we also repeated the analysis in the EdgeR package, using a recently published normalization method gene length corrected trimmed mean of M-values (GeTMM) (Smid et al., 2018). Both methods yielded very similar results with minor discrepancies (Figure 1—figure supplement 1C), and we chose to perform further downstream analysis using the DEseq-generated DE transcript list. A summary of transcriptomics results for core metabolic pathways can be found in Figure 1—figure supplement 2, and the list of transcripts that passed our DE cutoffs can be found in Supplementary file 1. The number of transcripts passing our DE cutoffs (6247) was extremely high, suggesting that multiple pathways are differentially regulated between the two species. We then performed gene set enrichment analysis (GSEA), searching against the Gene Ontology Biological Process (GO BP) database. Supplementary files 2 and 3 contain the list of enriched gene sets in PaLung and WI-38 cells respectively. Since the P. alecto genome is less fully annotated than the human genome, pathways with incomplete annotation may be incorrectly predicted to be downregulated in PaLung cells. Hence, we only studied differentially regulated pathways that were upregulated in PaLung. When we filtered PaLung-upregulated gene sets for significance (indicated by FDR < 0.25 and normalized enrichment score |NES| > 1), and for relevance to metabolism, only 21 gene sets remained. Many were relevant to secondary metabolism or anabolic/catabolic housekeeping, five were related to the TCA cycle and electron transport (including ‘ATP synthesis by chemiosmotic coupling, respiratory electron transport, and heat production by uncoupling proteins’) and three were related to hypoxic stress (such as ‘cellular response to hypoxia’ in the Reactome Pathway database). The genes belonging to both oxidative phosphorylation (OxPhos) (Figure 1C) and response to hypoxia (Figure 1D) gene sets had increased transcriptional expression in PaLung cells. This was interesting because, conventionally, these two pathways are active under opposing conditions (high oxygen for OxPhos vs low oxygen for hypoxia), and so would not be expected to vary in concert with each other. This hinted toward non-trivial regulation of central metabolism in PaLung cells.

Figure 1. RNAseq data analysis of PaLung and WI-38 cells for differential expression and pathway enrichment.

(A) Workflow of bioinformatics analysis pipeline for RNAseq data from PaLung (P. alecto) and WI-38 (H. sapiens) cells (n=3). (B) Heatmap showing the expression patterns for genes that passed our differential expression thresholds in the three WI-38 samples (W1–W3) and the three PaLung samples (P1–P3). (C and D) Gene set enrichment analysis (GSEA) identifies respiratory electron transport and cellular response to hypoxia as top metabolic pathways that are differentially regulated between PaLung and WI-38 cells. Shown here are the enrichment score plots for (C) respiratory electron transport and (D) cellular response to hypoxia.

Figure 1.

Figure 1—figure supplement 1. Analysis of transcriptomics data from PaLung and WI-38 cells.

Figure 1—figure supplement 1.

(A) Principal component analysis plots showing the separation of the PaLung and WI-38 transcriptomics datasets. (B) Volcano plot showing the genes that passed our differential expression (DE) thresholds (|log2 fold change [LFC]| ≥ 1 and false discovery rate (FDR) < 0.05). (C) Correlation between the gene fold changes estimated via the DESeq2 pipeline and the gene length corrected trimmed mean of M-values (GeTMM) pipeline (see Materials and methods). The points on the X-axis represent LFC of genes that passed our DE thresholds in DESeq2 but not in GeTMM. Conversely, points on the Y-axis represent genes that passed our DE thresholds in the GeTMM pipeline but not in the DESeq2 pipeline. (D) Pie chart showing that the gene set enrichment analysis (GSEA)-suggested upregulation of electron transport chain (ETC) in transcriptomics data is a result dominated by genes corresponding to Complex I subunits. C1=Complex I; C2=Complex II; C3=Complex III; C4=Complex IV of the ETC.
Figure 1—figure supplement 2. A summary of transcriptomics log fold changes (LFC) overlaid onto key metabolic reactions from central carbon metabolism.

Figure 1—figure supplement 2.

Shown here is a network of reactions in central carbon metabolism that includes reactions from multiple pathways (glycolysis, pentose phosphate pathway, TCA cycle, fatty acid oxidation, electron transport chain, urea cycle, and malate aspartate shuttle). The nodes of the network indicate metabolites and are shown as purple boxes with black outlines. The black arrows represent metabolic reactions. The colored boxes adjacent to the arrows represent the LFC of a gene involved in catalyzing the specific metabolic reaction (subunit/isoform/alternate genes). The number of boxes corresponds to the total number of genes participating in the reaction. The color scale indicates the extent to which the gene is upregulated in PaLung (blue) or WI-38 (red). Gray squares indicate that a particular gene/transcript implicated in the reaction was either not detected or did not pass our differential expression thresholds between PaLung and WI-38 samples.

Mitochondrial proteomics suggests that OxPhos is higher in PaLung cells compared to WI-38 cells

To test whether the whole-cell RNA differences were also reflected in mitochondrial composition, we performed tandem mass tag-based proteomic profiling in the mitochondrial fractions of PaLung and WI-38 cells. Profiling detected a total of 1469 proteins, that had peptides detected with high confidence in both species. There were no peptides detected in our experiment that were exclusively detected in high confidence in only one organism. Analysis using the gene ontology tool Enrichr confirmed that a majority of these proteins were likely obtained from a mitochondrial compartment (Figure 2—figure supplement 1A). When we performed DE analysis on the 1469 proteins (see Materials and methods, Figure 2A), 405 were differentially expressed between WI-38 and PaLung cells (Figure 2—figure supplement 1B). Of these 405 proteins, we identified 127 to be core mitochondrial proteins (as defined by MitoCarta and IMPI datasets), that were differentially expressed between WI-38 and PaLung cells (Supplementary file 4). Figure 2B shows the heatmap of row-normalized abundances of the 127 DE proteins. We observed that most of these 127 DE mitochondrial proteins are upregulated in bat PaLung samples (109), with very few downregulated proteins (18), suggesting increased mitochondrial activity in PaLung cells.

Figure 2. Proteomic data analysis of mitochondrial fractions of PaLung and WI-38 cells for differential expression, pathway enrichment, and electron transport chain (ETC) activity.

(A) Workflow of bioinformatics analysis pipeline for proteomics data from PaLung (P. alecto) and WI-38 (H. sapiens) cells (n=3). (B) Heatmap showing the expression patterns for the 129 differentially expressed mitochondrial proteins in the three WI-38 samples (W1–W3) and the three PaLung samples (P1–P3). (C) Differentially expressed mitochondrial proteins (nodes colored by log fold change) are overlaid on a network of mitochondrial protein-protein interactions (obtained from STRING) (W=WI-38 cells, P=PaLung cells). The nodes are then clustered with respect to reactome-annotated pathways. (D and E) Gene set enrichment analysis (GSEA) identifies citric acid cycle, oxidative phosphorylation, and Complex I biogenesis as top metabolic pathways that are differentially regulated between PaLung and WI-38 cells. Shown here are the enrichment score plots for (D) citric acid cycle and oxidative phosphorylation and (E) Complex I biogenesis. (F) Oxygen consumption rate (OCR) measurement of PaLung cells (blue) and WI-38 cells (red) plotted as mean ± SD from n>15 independent experiments. O=oligomycin, F=FCCP, R+A = rotenone+antimycin A. (G) Pie chart showing that the proteomic upregulation of the ETC, implied by GSEA, is dominated by genes for subunits of Complex I. C1=Complex I; C3=Complex III; C4=Complex IV of the ETC.

Figure 2.

Figure 2—figure supplement 1. Proteomic analysis of PaLung and WI-38 data.

Figure 2—figure supplement 1.

(A) Gene ontology category cellular (GO CC) enrichment using the Enrichr tool for compartmental enrichment on the 1469 genes detected in the proteomics dataset shows enrichment for the mitochondrial compartment. (B) Volcano plot showing the differentially expressed (DE) proteins (|log2 fold change| ≥ 1 and false discovery rate (FDR) < 0.05). Gray dots represent non-mitochondrial proteins that are not differentially expressed. Black dots show non-mitochondrial proteins that are differentially expressed. Red dots show all mitochondrial proteins.

To identify highly connected subnetworks that have strong expression changes, the set of 127 DE proteins was overlaid on a background of all known mitochondrial protein-protein interactions using STRING (with background obtained from MitoCarta and IMPI datasets). The results were clustered, and reactome pathway enrichment analysis of the resulting clusters showed enrichment for the following pathways: (1) TCA cycle and BCAA metabolism, (2) ETC, (3) fatty acid metabolism, and (4) protein import into mitochondria (Figure 2C).

We also performed GSEA using the abundances of the 1469 detected proteins as input against gene sets in the GO BP database. Supplementary files 5–6 show the enriched gene sets in PaLung cells and WI-38 cells, respectively. Supplementary file 5 shows that PaLung mitochondria were enriched for proteins in respiratory electron transport compared to WI-38 cells (Figure 2D). In particular, the gene set for Complex I biogenesis (Figure 2E) was significantly enriched in PaLung cells. For robustness, we also performed additional GSEA of the mitochondrial proteomics data under different conditions, e.g., leaving out outlier samples P1, W1 (Appendix 1—table 1), or when using a mitochondria-specific gene set list instead of the full GO BP gene set list (Appendix 2—table 1). Results from these additional analyses agree with our primary GSEA that OxPhos and Complex I proteins are enriched in PaLung mitochondria compared to WI-38 mitochondria. Proteomics thus agrees with transcriptomics that OxPhos genes are more highly expressed in PaLung cells than WI-38 cells, however, there were no indications of hypoxia-related gene sets being differentially regulated in the proteomics dataset. Hence, we decided to pursue experimental studies of OxPhos rates in PaLung and WI-38 cells.

PaLung cells have a lower OCR than WI-38 cells

To better characterize differences in OxPhos between PaLung and WI-38 cells, we monitored the oxygen consumption of these two cell lines using Seahorse XF Analyzer. Unexpectedly, the oxygen consumption rate (OCR) under basal conditions was lower in PaLung cells compared to WI-38 cells. Notably, bat cells have much lower maximal respiratory capacity compared to WI-38 cells, indicated by FCCP treatment (Figure 2F). This suggested that PaLung cells are able to carry out less OxPhos than WI-38 cells. While this observation agrees with earlier studies showing mild depolarization in the mitochondria of bats (Vyssokikh et al., 2020), it is in sharp contrast with our observation of increased OxPhos machinery in PaLung cells, using transcriptomics and proteomics. Taking a closer look at the omics results, we observed that the GSEA-flagged upregulation in OxPhos was driven mostly by the upregulation of Complex I subunits, for both the proteomic and transcriptomic data (Figure 2G, Figure 1—figure supplement 1D). This led us to hypothesize that in the basal state, PaLung cells might have partial decoupling of Complex I from the ETC, meaning that electrons emerging from Complex I might not proceed through the entirety of the ETC. This partial decoupling of Complex I might result in lower overall ATP synthesis, creating continued demand for ATP production from other sources. This would be consistent with our transcriptomic finding that both OxPhos-related and hypoxia-related genes were upregulated in PaLung cells. To build a self-consistent interpretation of these paradoxical omics and functional datasets, we proceeded to perform computational modeling of mitochondrial metabolism.

Computational flux modeling suggests that Complex II of the ETC may run in reverse in PaLung cells

To understand the metabolic consequences of having higher Complex I activity but lower overall respiration, we turned to constraint-based flux modeling (Figure 3—figure supplement 1). We started with Mitocore, a published model of mitochondrial metabolism, that includes both core mitochondrial reactions and supporting cytoplasmic reactions from central carbon metabolism (glycolysis, pentose phosphate pathway, folate cycle, urea cycle, etc.) (Smith et al., 2017). Mitocore provides the set of metabolic reactions that can occur in a cell without specifying the activity, abundance, expression, or utilization of each element. To establish a model for each species, we took a species-specific subset of Mitocore reactions (called a context-specific reconstruction) based on the presence and absence of gene/protein expression in the transcriptomic/proteomic data for each species (see Materials and methods). This resulted in a PaLung model with 409 reactions and 324 metabolites (Supplementary file 7), and a WI-38 model with 437 reactions and 341 metabolites (Supplementary file 8) as shown in Figure 3A. We then performed uniform flux sampling (5000 samples) of the two models without imposing any constraints on either model (Figure 3B). Simulations resulted in a feasible flux distribution for each reaction in both the PaLung model and the WI-38 model, which can be depicted as a frequency histogram. In each chart in Figure 3B and C, the X-axis represents the metabolic flux value for the corresponding reaction, and Y-axis represents the frequency/probability of the reaction having the specific flux value. Figure 3B shows these histograms for the fluxes through Complex I–III in both the PaLung and WI-38 metabolic models, under unconstrained conditions (control simulation).

Figure 3. Metabolomic data and model-based analysis of mitochondrial metabolism in PaLung and WI-38 cells.

(A) Schematic showing the metabolic modeling pipeline. We begin with the context-specific reconstruction of a metabolic model – the process where a generic mitochondrial model (Smith et al., 2017) is tailored specifically to PaLung and WI-38 cells using proteomic and transcriptomic expression patterns. The individual metabolic models are then simulated using constraint-based flux sampling methods to give a distribution of possible fluxes for each reaction in the model. Comparing these flux distributions between the two models allows the detection of metabolic reactions that are likely to be differentially regulated in response to user-imposed constraints on metabolism. Simulations are performed on both the PaLung and WI-38 models under no constraints or with constraints on Complex I and mitochondrial O2 intake. (B) Sample histograms showing the feasible flux distributions for electron transport chain (ETC) reactions in the unconstrained PaLung (P) and WI-38 (W) metabolic models. (C) Flux distributions of ETC reactions in PaLung (P) and WI-38 (W) cells when PaLung cells are constrained to have higher flux through Complex I of ETC but lower oxygen intake in the mitochondria. (D) Heatmap showing differentially regulated metabolites from central carbon metabolism in the three WI-38 (W1–W3) and the three PaLung (P1–P3) samples. (E) Intra-sample ratios of metabolites from the TCA cycle in either PaLung (P) or WI-38 (W) cells, plotted as mean ± SD from three independent experiments. * and *** represent p-value≤0.05 or ≤0.001 respectively (unpaired Student’s two-sided t-test with Benjamini-Hochberg correction for multiple hypothesis testing).

Figure 3.

Figure 3—figure supplement 1. Schematic of the workflow for metabolic flux modeling.

Figure 3—figure supplement 1.

(A) Left is a toy network of metabolic reactions showing a few reactions from the glycolysis pathway for illustration. This network of metabolic reactions (defined by nodes and arrows) is converted from the format of reaction equations into the format of a matrix called the stoichiometric matrix. Each row of the matrix represents a single metabolite in the network and each column represents a single reaction in the network. The (i,j) element of matrix S represents the stoichiometric coefficient of metabolite i in reaction j. Stoichiometric coefficients indicate how many copies of each reactant or product are consumed or produced when the reaction is utilized (often 1 or –1). By converting all the reactions of a metabolic network into this matrix format, a single column represents a single reaction in the metabolic network, including all metabolites produced or consumed in it. For example, the first column of the matrix (corresponding to reaction R1) has the value of –1 for the Glc row and 1 for G6P row, indicating that in reaction R1, one unit of Glc is consumed and one unit of G6P is produced. Likewise, each row of the matrix represents all sources of production and consumption for a single metabolite. For example, the second row of the matrix (corresponding to the metabolite G6P) has the values of 1, –1, and –1 for the reactions R1, R2, and R3, respectively. This indicates that one unit of G6P is produced in reaction R1, while one unit of G6P is consumed in reactions R2 and R3, respectively. (B) Example of a flux vector. A flux vector contains as many elements as the number of reactions in the network, and their values are typically unknown. Flux values indicate the rate at which the corresponding reaction is utilized (a forward reaction is a positive flux, and a reverse reaction is a negative flux). The goal of flux sampling is to obtain values (or distributions) for each element of the flux vector. (C) The product of the stoichiometric matrix and the flux vector (written as the product Sv) results in a set of linear algebra equations. Each linear equation has flux variables and stoichiometric coefficients. If we assume the flux is balanced in the system, then everything produced has somewhere to go or some reaction to consume it. This steady-state assumption is described mathematically by setting the matrix product to zero (Sv = 0). (D) The process of flux sampling searches the space of possible flux vectors v to find those that satisfy Sv = 0. Sampling is typically necessary because the set of equations from (C) is usually underdetermined and does not have a unique solution for the flux variables v. The computational sampling process outputs flux vectors that satisfy the equations (i.e. feasible vectors), and that also capture a diversity of feasible behaviors for how the metabolic network could achieve steady state. From this set of flux vectors, we plot histograms (distributions) of the feasible values of the flux variable for each reaction in the metabolic network. This theoretical delineation of feasible behaviors can be used for drawing biological inferences by assuming that the actual biological pathway activity falls within the feasible range of the theoretical model. When comparing two models such as bat and human, inferences arise from reactions or pathways where feasible behaviors become infeasible or vice versa. This can be seen by identifying reactions, whose flux histograms show large differences between the human and bat models.
Figure 3—figure supplement 2. Analysis of the metabolomics data from PaLung and WI-38 cells.

Figure 3—figure supplement 2.

(A) Principal component analysis (PCA) plot shows the separation between the three PaLung samples and the three WI-38 samples in the metabolomics data. (B–D) Bar plots of AMP (B), the ratios of Glu/Gln (C), and Asp/Asn (D) in PaLung cells (P, blue) and WI-38 cells (W, red) are shown. Bars are the mean ± SD from three independent experiments (n=3). ** or *** represents p-value≤0.01 or ≤0.001 respectively (unpaired Student’s two-sided t-test). Note that PaLung cells have a higher glutamate-to-glutamine ratio and a lower aspartate-to-asparagine ratio than WI-38 cells. This is also supported by the high expression of the gene ASNS (asparagine synthetase) in PaLung cells (Supplementary file 1). ASNS is a cytoplasmic enzyme that can convert aspartate to asparagine concomitantly with the hydrolysis of glutamine to glutamate, allowing for glutamate to be used as an alternative source for energy, or as a precursor metabolite for the synthesis of glutathione.

We next imposed the following two constraints, which are hypothetical differences between bat and human cells, suggested by the above transcriptomic/proteomic and oxygen consumption measurements. The first constraint is that the PaLung model must have greater activity of Complex I than the WI-38 model. This constraint was inspired by the transcriptomic/proteomic data (Figures 1C and 2F) and by inferring that the flux through Complex I in PaLung cells would be greater than the flux through Complex I in WI-38 cells. (see ‘Constraints for flux sampling’ in Materials and methods). The second constraint, based on our oxygen consumption data, is that the WI-38 model must have higher oxygen intake into the mitochondria than the PaLung model. When the PaLung and WI-38 models were simulated under these two constraints, the flux histograms for PaLung cells had shifted to very low or negative flux values for Complex II (Supplementary file 9). Complex II is also called succinate dehydrogenase (SDH) and is part of both the TCA cycle and the ETC. SDH generally catalyzes the conversion of succinate into fumarate, accompanied by a reduction of the endogenous Quinone pool. However, SDH has also been documented to catalyze the reverse reaction, converting fumarate to succinate, although this is unconventional. Such an unconventional ETC paradigm where Complex I proceeds forward but Complex II proceeds in reverse (toward the accumulation of succinate) has been observed in other systems, e.g., in murine retinal tissues (Bisbach et al., 2020) and human hearts (Chouchani et al., 2014), under ischemic or hypoxic conditions. In these cases, the resulting accumulation of succinate was found to be useful in fueling other cells in the tissue or in avoiding reperfusion injury when ischemic conditions were abruptly removed. During conventional ETC, both Complex I and Complex II operate in parallel and produce electrons that are shuttled downstream to Complex III, Complex IV, and ATP synthase. However, prior work (Bisbach et al., 2020; Chouchani et al., 2014) has documented an alternative in which the electrons obtained from Complex I can be consumed by Complex II operating in reverse, rather than traversing the rest of the ETC. Accordingly, the low or negative flux values for Complex II in our PaLung simulations indicate that the electrons obtained from Complex I may accumulate at Complex II or potentially even get consumed by Complex II operating in reverse (bypassing the rest of the ETC) in PaLung cells. To further interrogate central carbon metabolism in PaLung cells and to validate if our predictions of unconventional SDH activity might be borne out by metabolic measurements, we undertook targeted metabolomics of small organic compounds for both PaLung and WI-38 cells.

Metabolomics supports the possibility of low/reverse Complex II activity in PaLung cells

Mass spectrometry-based targeted metabolomics was performed on both PaLung and WI-38 cell lines, resulting in the absolute quantification of 116 metabolites from central carbon metabolism (Figure 3D, Figure 3—figure supplement 2, Supplementary file 10). Since we were comparing different species, we looked at relative ratios of metabolites within each species. As predicted by our hypothesis, metabolomics showed that the ratio of succinate-to-fumarate was much higher in PaLung cells (5.44±0.67) compared to WI-38 cells (1.07±0.012) (p<0.001), consistent with succinate accumulation in PaLung cells (Figure 3E). In contrast, the ratio of other serial TCA metabolites, e.g., citrate/succinate or fumarate/malate, showed only mild or insignificant differences between PaLung and WI-38 cells. We interpreted this as an indication that the TCA cycle acts in a truncated manner in PaLung cells, with SDH operating in reverse to support succinate accumulation. Interestingly, previous studies have documented such SDH phenomena (with the low or reverse activity of Complex II) during ischemic/hypoxic states that had low metabolic rates and high levels of AMP (Chouchani et al., 2014; Bisbach et al., 2020). This led us to wonder if PaLung cell metabolism might resemble an ischemic-like state despite oxygen and nutrient availability, in which case we would expect to see low metabolic rates and high AMP levels in PaLung cells.

PaLung cells exhibit basal metabolism that resembles an ischemic-like state and can tolerate glucose deprivation better than WI-38 cells

To further understand the consequences of the truncated TCA cycle and its implications on an ischemic-like state, we looked at the abundances of other metabolites in metabolomics datasets. Looking at metabolites from glycolysis, pentose phosphate pathway, TCA cycle, and the levels of amino acids, we found that most metabolites were at much lower levels in PaLung cells compared to WI-38 cells, except for a small subset of metabolites in glycolysis (Figure 4A–C). This suggested an overall slower metabolic turnover in PaLung cells. PaLung cells also had a much higher proportion of AMP (Figure 4D), which would be expected in an ischemic setting.

Figure 4. PaLung cells show basal metabolism that resembles an ischemic-like state.

Figure 4.

(A) Fold changes (mean PaLung/mean WI-38) of metabolite abundances, obtained through targeted metabolomics profiling of central carbon metabolites in PaLung and WI-38 cells. (B and C) Amino acid and metabolite changes in TCA cycle (PaLung/WI-38) in PaLung cells and WI-38 cells. (D) Bar plots of AMP/(AMP+ADP+ATP), ATP/ADP ratio, and lactate amounts in PaLung (P) and WI-38 (W) cells plotted as mean ± SD from three independent experiments. ** and *** represent p-value≤0.01 or ≤0.001 respectively (unpaired Student’s two-sided t-test with Benjamini-Hochberg correction for multiple hypothesis testing). (E) Extracellular acidification rate (ECAR) measurement of PaLung (blue) and WI-38 (red) cells plotted as mean ± SD from n>15 independent experiments (2-DG=2-deoxy-D-glucose). (F) Phase contrast images of PaLung and WI-38 cells with or without glucose deprivation for 96 hr. Scale bar, 100 µm.

Furthermore, our metabolomics dataset revealed that PaLung cells had a threefold lower ATP/ADP ratio compared to WI-38 cells (p<0.001). ATP/ADP ratio is a golden standard for the measurement of cellular energy status, and our results suggest lower metabolism in PaLung cells. Another corroborating measurement for energy status is adenylate energy charge, with lower values indicating slower metabolism (Atkinson and Walton, 1967). We found that PaLung cells had lower adenylate energy charge compared to WI-38 cells (p<0.001) (Supplementary file 10). From these parameters, we infer that PaLung cells have lower ATP synthesis, consistent with our earlier observations of lower OCR in PaLung cells (Figure 2F). To check if slower metabolism would also translate to slower glycolysis in PaLung cells, we first checked our transcriptomics and metabolomics data. Both omics datasets showed mixed signals along the glycolysis pathway (with partial upregulation and partial downregulation) (Figure 4A and Figure 1—figure supplement 2), yielding no predictions for glycolysis utilization. However, metabolomics showed higher levels of intracellular lactate in WI-38 cells than in PaLung cells, suggesting that PaLung cells could have lower glycolytic flux than WI-38 cells. Finally, we assessed glycolysis levels using a Seahorse XF Analyzer to quantify the extracellular acidification rate (ECAR). This revealed that PaLung cells had lower ECAR compared to WI-38 cells (Figure 4E). From these observations of low ECAR, low lactate, and low ATP/ADP ratio, combined with the earlier findings of low OCR, we conclude that PaLung cells have less energy production and a lower level of basal metabolic activity than WI-38 cells.

Reduced OxPhos and glycolysis in PaLung cells somewhat resemble an ischemic-like state with inadequate oxygen and glucose supply. Therefore, PaLung cells may be resistant to metabolic stress. To test this, we subjected PaLung and WI-38 cells to glucose deprivation for 96 hr. Interestingly, PaLung cells displayed higher viability than WI-38 cells after glucose deprivation (Figure 4F), despite starting with lower stores of internal energy (e.g. glucose-6-phosphate, acetyl-CoA, and lower adenylate charge). These data suggest that PaLung cells do have a better ability to tolerate metabolic stress compared with WI-38 cells, consistent with low basal metabolic activity.

PaLung cells have higher ROS compared to WI-38 cells but lower expression of many antioxidant genes

Previous studies of ischemic metabolism, in which SDH functions in reverse, have showed that high levels of ROS were generated as a result. To test if the same would be true in PaLung cells, we performed the MitoSOX assay to measure superoxide levels in the mitochondria of PaLung cells and WI-38 cells. Indeed, PaLung cells showed higher mitochondrial superoxide levels compared to WI-38 cells, both under basal conditions and when ETC was inhibited using antimycin A (Figure 5A). Searching our transcriptomic data for genes involved in the redox control, we observed that there were significant differences in the expression of redox control genes in PaLung and WI-38 cells (Figure 5B). Notably, SOD1 and SOD2, key enzymes that convert mitochondrial superoxide to the more toxic intracellular ROS, are less expressed in PaLung cells than in WI-38 cells, consistent with higher mitochondrial superoxide levels in PaLung cells. Interestingly, glutathione peroxidase 3 (GPX3), a well-known antioxidant enzyme that reduces hydrogen peroxide or organic hydroperoxides using glutathione, was found to be highly upregulated in PaLung cells compared to WI-38 cells. This led us to test levels of glutathione, a non-enzyme antioxidant for ROS detoxification, in the two cell lines.

Figure 5. Reactive oxygen species (ROS) and antioxidant system measurements in PaLung and WI-38 cells.

(A) MitoSOX measurement of PaLung and WI-38 cells with or without antimycin A treatment for 1 hr. Antimycin A is an electron transport chain (ETC) inhibitor known to induce superoxide generation. (B) Bar charts showing the expression levels of antioxidant genes that passed our differential expression thresholds (as transcripts per million [TPM]) in PaLung (blue) and WI-38 (red) cells. Genes have been sorted in increasing order of P/W fold change. (C) Bar plots show the ratio of reduced to oxidized glutathione (GSH/GSSG), total glutathione normalized to protein content (GSH+GSSG), and the ratio of NADPH/NADP in PaLung (P) and WI-38 (W) cells. For all panels, bars are the mean ± SD from three independent experiments (n=3). *, **, or *** represents p-value<0.05, ≤0.01, or ≤0.001 respectively (unpaired Student’s two-sided t-test with Benjamini-Hochberg correction for multiple hypothesis testing).

Figure 5.

Figure 5—figure supplement 1. Increased phosphorylation of NAD cofactors in PaLung cells.

Figure 5—figure supplement 1.

(A) Ratio of NADP/NAD in PaLung (P) and WI-38 (W) cells. (B) Bar graphs show the expression levels of the genes involved in NAD synthesis and phosphorylation (as transcripts per million [TPM]) in PaLung (blue) and WI-38 (red) cells. *, **, or *** represents p-value≤0.05, ≤0.01, or ≤0.001 respectively (unpaired Student’s two-sided t-test with Benjamini-Hochberg correction for multiple hypothesis testing).

PaLung cells have a robust glutathione NADPH system to counter intracellular ROS

To test if the glutathione system may help PaLung cells tolerate intracellular ROS, we measured intracellular glutathione and NADP(H) levels in both PaLung and WI-38 cells. The ratio of reduced to oxidized glutathione (GSH/GSSG ratio) was not significantly different between the two cell lines (WI-38 cells: 36.45±3.6 vs PaLung cells: 30.94±2.97). However, PaLung cells had a twofold higher concentration of total intracellular glutathione, compared to WI-38 cells (Figure 5C). We also found that PaLung cells had a higher NADPH/NADP ratio (1.5-fold) compared to WI-38 cells. Overall, the metabolomics data indicated that PaLung cells had a nearly 2.5-fold higher ratio of NADP/NAD, i.e., a higher resting concentration of phosphorylated to unphosphorylated NAD (Figure 5—figure supplement 1A). This was also supported by transcriptomics which showed an upregulation of NADK (NAD kinase) and supporting enzymes required to synthesize and phosphorylate NAD (Figure 5—figure supplement 1B). Taken together these results suggest that PaLung cells maintain a higher standing pool of glutathione and a higher NADPH concentration to counter ROS generated due to ischemic-like metabolism.

PaLung cells are resistant to ferroptosis

Many earlier studies have shown that ischemic conditions can induce cell death via ferroptosis, which also depends on accumulated ROS and low glutathione levels (Chen et al., 2021). We wondered if apart from combatting ROS, the high glutathione levels might also help PaLung cells avoid ischemia-induced ferroptosis. Recent studies have reported that ischemia-induced ferroptosis causes tissue damage and that inhibition of ferroptosis attenuates ischemia-induced cell death (Xie et al., 2019; Liao et al., 2021). Ferroptosis is a non-apoptotic, programmed form of cell death, which is iron-dependent and occurs via glutathione depletion-induced lipid peroxidation (Ursini and Maiorino, 2020). Given that PaLung cells showed upregulation of glutathione/NADPH antioxidant system and genes related to hypoxia response (Figure 1D) and ischemic metabolism, we tested whether PaLung cells can better tolerate ferroptosis-inducing conditions. Ferroptosis was induced by erastin, an SLC7A11/xCT inhibitor. While WI-38 cells showed high sensitivity to erastin-induced ferroptosis that were prevented by a ferroptosis inhibitor ferrostatin-1, PaLung cells were resistant to erastin-induced ferroptosis (Figure 6A and B). Similarly, PaLung cells were more resistant to cystine deprivation-induced ferroptosis, compared to WI-38 cells and had almost sevenfold lower cell death than WI-38 cells (3.5% cell death in PaLung cells compared to 24.4% cell death in WI-38 cells upon cystine deprivation) (Figure 6C and D). These results suggest that PaLung cells are more resistant to ferroptosis compared to WI-38 cells.

Figure 6. PaLung cells display high resistance to ferroptosis.

Figure 6.

(A and B) WI-38 or PaLung cells were treated with 2.5 µM erastin and/or 1 µM ferrostatin-1. Representative images were taken at 6 hr using phase contrast microscopy (A). Propidium iodide (PI) exclusion assay was performed 24 hr after erastin treatment (B). (C and D) WI-38 or PaLung cells were cultured in media with or without cystine. Ferrostatin-1 (1 µM) was treated simultaneously. Representative images were taken at 8 hr using phase contrast microscopy (C). PI exclusion assay was performed at 24 hr after cystine deprivation (D). (E–H) WI-38 or PaLung cells were cultured in media with or without cystine for 6 hr. Intracellular glutathione levels were measured. Reduced glutathione (GSH)/oxidized glutathione (GSSG) ratio (E), total glutathione (the sum of GSH and GSSG) (F), GSH (G), and GSSG (H) levels were measured. Scale bars, 50 μm. The mean ± SD of three independent experiments is shown.

To better understand the link between glutathione concentrations and ferroptosis resistance in PaLung cells, we measured the amounts of glutathione in WI-38 and PaLung cells under cystine deprivation. Both WI-38 and PaLung cells showed decreased GSH/GSSG ratio under cystine deprivation, which indicates both cell lines are under oxidative stress (Figure 6E). Cystine deprivation decreased total glutathione levels in WI-38 cells by 94.5%, nearly depleting glutathione reserves. In contrast, cystine deprivation in PaLung cells resulted in only a modest 29.5% decrease, and PaLung cells still maintained a total glutathione level higher than non-cystine-deprived WI-38 cells. The higher levels of total glutathione, GSH, and GSSG under cystine deprivation might explain the high resistance of PaLung cells under ferroptosis-inducing conditions (Figure 6F, G, and H).

Discussion

In this study, we integrated high-throughput omics and computational metabolic modeling to perform a novel comparison of central metabolism between cell lines from two mammalian species. Specifically, we have identified core differences in mitochondrial metabolism between primary lung fibroblasts of the black flying fox fruit bat P. alecto (PaLung) and the primary human lung fibroblast cell line WI-38. Although data are still limited to these two cell lines, this is the first comprehensive analysis combining proteomics, transcriptomics, metabolomics, and constraint-based flux modeling between human and bat cells. Our analysis suggests that PaLung cells exhibit basal metabolism that resembles an ischemic-like state and that this state may be linked to low or reverse activity of Complex II of the ETC (also called SDH). Compared to WI-38 cells, PaLung cells also show a higher tolerance to cellular stresses such as nutrient deprivation and ischemia/ROS-driven cell death via ferroptosis.

Hypoxia response (including glycolysis) and OxPhos compensate for each other by gene expression, dependent on different levels of oxygen availability. However, our analyses (GSEA from whole-cell transcriptomics) indicated that genes related to glycolysis and OxPhos were simultaneously upregulated in PaLung cells (bat) compared to WI-38 cells (human). Although multiple glycolytic genes were upregulated in PaLung cells (Figure 1—figure supplement 2), lactate production (inferred via both LDH expression levels and ECAR measurement) in PaLung cells was not higher than in WI-38 cells. This raises an important caveat about using ECAR to infer glycolytic flux, because ECAR depends only on the amount of lactate produced, and glycolytic pyruvate that enters the TCA cycle may remain invisible to ECAR measurements.

Our observation about the upregulation of OxPhos in PaLung cells via GSEA (transcriptomics and proteomics) also raises an important point of concern, namely that high signals of a certain pathway in GSEA can be driven by a narrower subset of genes. For instance, in both our transcriptomic and proteomic GSEA results, the ETC genes upregulated in the PaLung samples did follow the frequency distribution expected when taking into account the total subunit counts of each ETC complex (CI: 44, CII: 4, CIII: 10, and CIV: 19 subunits). However, the sheer absolute number of Complex I genes upregulated was enough for GSEA to declare the entire OxPhos gene set as upregulated in PaLung cells. While this does not say if the other ETC complexes were upregulated in PaLung, it shows the outsized effect Complex I genes have on the entire OxPhos pathway, simply because of overwhelming numbers. Hence, while interpreting GSEA results, it is necessary to survey the constituent genes that are responsible for the predicted up/downregulation of the overall pathway. Indeed, we found that the upregulation of OxPhos pathway in PaLung cells in GSEA was due to high gene expression of the Complex I components in the ETC. However, our experimental analysis revealed a puzzling result: despite high Complex I expression, oxygen consumption was low in PaLung cells.

To interpret the metabolic implications of heightened Complex I and lower oxygen consumption, we turned to constrained-based metabolic flux sampling. Flux sampling is a technique that can simulate possible states of a metabolic network. Compared to higher resolution methods such as isotope-labeled fluxomics, flux sampling is a coarse-grained qualitative approach to study metabolic flux. However, it can be used as a tool to generate testable hypotheses in exploratory studies such as ours. Indeed, metabolomics was able to verify our simulation predictions of succinate accumulation in PaLung cells, a phenomenon that could result from low/reverse activity (fumarate to succinate) of Complex II of ETC (SDH). The succinate-to-fumarate ratio in well-oxygenated PaLung cells was found to be similar to those found in ischemic states of human cells from multiple tissues (Chouchani et al., 2014). From this, we infer that the TCA cycle in PaLung cells does something different than the TCA cycle in human WI-38 cells, which is also corroborated by our observation that PaLung cells have a low metabolic rate. We also confirmed using cell culture assays that PaLung cells have a much longer survival during glucose deprivation than WI-38 cells.

We also found that PaLung cells had higher loads of mitochondrial ROS, consistent with having the lower expression of SOD1 and SOD2. On the other hand, PaLung cells highly express glutathione peroxidase 3 (GPX3), contain twofold higher levels of glutathione, and maintain a higher NADPH/NADP ratio compared to WI-38 cells. These data suggest that PaLung cells maintain a standing pool of NADPH and glutathione that can act against high levels of mitochondrial ROS. Therefore, the previously published generalization that bats have lower free radical production than other mammals (Brunet-Rossinni, 2004; Brown et al., 2009) may require moderation. Future work can address whether PaLung cells have a higher threshold for homeostatic ROS and whether ROS have both beneficial and detrimental effects in bats, as has been observed in other organisms (Clément and Pervaiz, 2001; Shields et al., 2021).

The key novelty of our results is the suggestion of a basal ischemic-like state in PaLung cells accompanied by higher ROS production and higher glutathione and NADPH levels. Multiple studies have shown strong links between ischemia, ROS, and ferroptosis (an iron-dependent, non-apoptotic form of cell death). Despite having ischemic-like metabolism and higher ROS levels, we observed that PaLung cells were more resistant to ferroptosis than WI-38 cells. Ferroptosis occurs via ROS-induced lipid peroxidation following ischemia and is conventionally blocked via the action of glutathione peroxidase 4 (GPX4). While GPX4 expression levels were low in PaLung cells, GPX3 (a commonly secreted form of glutathione peroxidase) was highly expressed. Future studies are needed to determine the function of GPX3 in the stress response and whether it contributes to the ferroptosis resistance of PaLung cells.

Our identification of a metabolic state in bats that resembles ischemia is also tied to conventional wisdom about bat flight. During flight, the metabolic rate of bats can go 2.5–3 times higher than that of other mammals, consuming approx. 1200 calories per hour, resulting in an immense drain on stored energy reserves, and the potential depletion of up to 50% of stored energy in fructivore bats (Thomas, 1975; Voigt and Speakman, 2007; Kelm et al., 2011). Thus, a basal state with a low metabolic rate may serve to conserve energy during non-flying periods. In addition, the accumulated succinate during this ischemic-like state can also be transported into the bloodstream and serve as a metabolic stimulant for other fast-metabolizing cell types as needed, similar to examples of succinate shuttling between two cell types in retinal tissues (Bisbach et al., 2020). Low metabolic rates have also been linked to longevity by many studies, which demonstrate that calorie restriction contributes to an extended lifespan across various species. In our experiments, P. alecto also showed higher expression of genes involved in NAD synthesis and its phosphorylation pathways, in agreement with other studies that have linked higher flux through the NAD biosynthesis pathway with slower aging (Chini et al., 2017).

Given that we compare two vastly different species, our work is subject to some technical limitations due to assumptions of comparability. Specifically, (1) we perform GSEA using human-derived GO terms and gene sets, (2) we map proteomic peptides between human and P. alecto proteins, and (3) we use human-derived metabolic mitochondrial models for computational flux sampling. However, such assumptions are unavoidable and currently serve as the best recourse, until further research into bats can provide better bioinformatic tools more suited to bats. In addition, reference genomes for less studied organisms like bats may not be as well annotated as other organisms. Since we perform transcriptomic analyses only using genes identified in both species, observed downregulation of a pathway in PaLung cells might be because the constituent genes were not annotated as well, and not because they weren’t expressed as highly. For exactly this reason, we have focused on pathways that were upregulated in bats, and we made no conclusion about pathways downregulated in bats.

A limitation of our study is that we performed multi-omics comparisons between individual cell lines. Extending the current study to primary tissues or cell lines from other species of bats might provide further information about generalized metabolic differences between bats and humans. The differences we observe in our study could also be cell-type dependent. For example, our choice of fibroblast cell lines may explain why we observed PaLung cells to have lower oxygen consumption than WI-38 cells, even under uncoupled conditions. Future work can study muscle cells in the same species to obtain broader metabolic insights. The current study can be considered a starting point to both generate hypotheses for future work and to establish analysis pipelines for inter-species comparisons of metabolism, using multi-omics data and computational flux modeling simulations.

In summary, using multi-omics datasets (transcriptomic, proteomic, metabolomic) and computational modeling, we have compared the basal metabolic states of a human cell line and a P. alecto cell line. Our work points toward important differences in ETC regulation (potential low/reverse Complex II activity) and antioxidant response (higher glutathione and NADPH) that could contribute to ischemia/redox management and ferroptosis resistance in P. alecto. Future research can extend the idea further to identify if such regulation is a pan-bat feature and if such metabolic fingerprints might also contribute to cancer resistance and increased longevity in bats.

Materials and methods

Cell cultures and reagents

WI-38 cells (RRID: CVCL_0579) from H. sapiens (RRID: NCBITaxon_9606) were purchased from Coriell Institute. PaLung, a lung-derived cell line from P. alecto (RRID: NCBITa xon_9402), was established as previously described (Koh et al., 2019; Crameri et al., 2009). In brief, PaLung cells were derived from the lung tissue of a single female P. alecto (n=1) and were established as a primary cell line through a process that combined trypsinization and physical disruption. This was followed by culturing in Dulbecco’s modified Eagle’s medium (DMEM)/F12-Hams (Sigma) medium, supplemented with 15% bovine calf serum (Hyclone), 100 units/ml penicillin, 100 µg/ml streptomycin, and 50 µg/ml gentamycin (Sigma) in 5% CO2-humidified atmosphere at 37°C. Upon their establishment, all the cell lines were cultured in high-glucose DMEM (#11965, Gibco, Life Technologies) supplemented with 10% FBS (HyClone, GE Healthcare Life Science), penicillin (100 units/ml), and streptomycin (100 mM/ml; Gibco, Life Technologies) in 5% CO2-humidified atmosphere at 37°C. WI-38 cells were authenticated by Coriell Institute. The PaLung cell line is a novel primary cell line derived from bats and was not authenticated since no standard workflow has been established for authentication of bat-derived cells. WI-38 and PaLung cell lines tested negative for mycoplasma contamination. For glucose deprivation, cells were washed with phosphate-buffered saline (PBS) three times and cultured in glucose-free DMEM (#10966, Gibco, Life Technologies) with 10% dialyzed FBS. For cystine deprivation, cells were washed with PBS and cultured in cystine, methionine, glutamine-free DMEM (#21013024, Gibco, Life Technologies) supplemented with 0.2 mM L-methionine, 4 mM L-glutamine, and 10% dialyzed FBS. L-glutamine was purchased from Invitrogen. L-methionine and L-cystine were kindly provided by Dr. Jean-Paul Kovalik (Duke-NUS Medical School, Singapore). Oligomycin, FCCP, rotenone, antimycin A, 2-deoxyglucose, and erastin were purchased from Sigma-Aldrich. Ferrostatin-1 was purchased from Med Chem Express.

PI exclusion assay

Cells were stained with PI to determine the percentage of cell death. Media containing floating cells were collected, combined with trypsinized cells, and centrifuged. The cell pellet was washed once with PBS. After centrifugation, cells were resuspended and stained with PI (10 μg/ml) for 10 min at room temperature. Data were collected with MACSQuant analyzer (Miltenyi Biotec). Quantification and analysis of the data were done with FlowJo software (RRID: SCR_008520).

Transcriptomics (sample preparation and initial bioinformatics)

Three independent sets of RNA were collected from WI-38 and PaLung cells at different passages (n=3). In brief, cells were subcultured at a 1:2 ratio and collected at passage numbers 26 (W1), 30 (W2), and 35 (W3) for WI-38 cells, and 5 (P1), 6 (P2), and 7 (P3) for PaLung cells. RNA was extracted from WI-38 and PaLung cells using RNeasy Plus Mini Kit (QIAGEN). 1000 ng of total RNA from each sample was used to generate RNAseq libraries using TruSeq Stranded Total RNA Library Prep Gold according to the manufacturer’s instructions (Illumina). Library fragment size was determined using DNA1000 Assay on the Agilent Bioanalyzer (Agilent Technologies). 2×150 PE sequencing was subsequently performed on the libraries using HiSeq3000 equipment (Illumina). The resulting reads were cleaned/trimmed and demultiplexed, followed by mapping to either the PA genome (annotation 102, 2018) or the hg19 genome using cufflinks/Tophat.

Transcriptomics, DEseq, GeTMM, and GSEA

For DE analysis, the raw read counts were input to the R package DESeq2 (Love et al., 2014). Genes with counts per million < 1 in more than three out of six samples (three from PaLung and three from WI-38) were discarded from downstream analysis. To account for differences in transcript length between the two species, the individual transcript lengths were supplied as an additional normalization factor to DESeq2. Since the inter-species analysis of RNAseq data does not have conventional workflows, we repeated DE gene identification using an alternative workflow with GeTMM (Smid et al., 2018). This method uses RPKM as input and hence accounts automatically for differences in transcript lengths. This workflow was implemented using the R package EdgeR (Robinson et al., 2010). For GSEA, we used the GSEApreranked module in GSEA version 4.1.0 (Subramanian et al., 2005), using the DESeq2 results as input. For ranking genes, we used the π-value metric [LFC*(-log10pvalue)] (LFC = log2 fold change) (Xiao et al., 2014). GSEA was run against the complete gene set list that includes GO BP with pathways from databases such as Reactome and KEGG (containing 18,356 gene sets), downloaded from https://download.baderlab.org/EM_Genesets.

Mitochondria isolation

Mitochondria were isolated from WI-38 and PaLung cells using mitochondrial isolation kit from Miltenyi Biotech as described in the manufacturer’s protocols. In brief, 1×107 cells from WI-38 and PaLung cells were lysed in the ice-cold hypotonic lysis buffer for 60 s followed by mechanical disruption of cell membrane using a mini homogenizer pestle gun for 60 s. The suspension was centrifuged at 700 × g for 5 min at 4°C and supernatant was collected. The supernatant containing the mitochondria was incubated with the TOM-22 antibody-conjugated with MACS magnetic beads (Miltenyi Biotech) and pulled down using the magnetic columns. The mitochondria fractions were eluted with 0.1 M glycine pH 3.5, neutralized with Tris-HCl pH 7.5, and stored at 80°C until the time of proteomics profiling.

Proteomics profiling and analysis

The mitochondrial fractions were lysed in 8 M urea pH 8.5 and incubated with 20 mM tris(2-carboxyethyl)phosphine hydrochloride for 20 min at 25°C followed by alkylation with 55 mM chloroacetamide at 25°C for 30 min. Proteins were digested with trypsin overnight at 25°C. Digested peptides were acidified with 1% trifluoroacetic acid, desalted on C18 plates Oasis (Waters), and labeled with TMT sixplex reagent (Thermo Scientific) according to the manufacturer’s protocol. Labeled samples were further fractionated with high pH reverse phase using spin columns packed in-house, and five fractions were collected: 10%, 17.5%, 25%, 30%, and 50%. The fractions were separated on a 50 cm × 75 µm Easy-Spray column using Easy-nLC system coupled with an Orbitrap Fusion Tribrid mass spectrometer (Thermo Scientific). The LC-MS/MS parameters for fusion: peptides were separated over a 120 min gradient, using mobile phase A (0.1% formic acid in water) and mobile phase B (0.1% formic acid in 99% acetonitrile), and eluted at a constant flow rate of 300 nl/min. Acquisition parameters were as follows: data-dependent acquisition with survey scan of 60,000 resolution, AGC target of 4×105, and maximum injection time (OT) of 100 ms; MS/MS collision induced dissociation in Orbitrap 15,000 resolution, AGC target of 1×105, and maximum IT of 120 ms; collision energy NCE = 35, isolation window 1.0 m/z. Peak lists were generated in Proteome Discoverer 2.1, and a search was done using Sequest HT (Thermo Scientific) with human Uniprot and fruit bat Uniprot databases. The following search parameters were used: 10 ppm MS; 0.06 Da for MS/MS with the following modifications: oxidation (M) deamidation (N,Q), TMT adduct (N-term, K) carbamidomethyl (C). Peptides detected in WI-38 samples were automatically mapped to their source human protein Uniprot ID. To map the peptides detected in PaLung samples to the corresponding P. alecto Uniprot ID, we used a two-step approach. First, we used inParanoid (O’Brien et al., 2005) to identify known orthologs between H. sapiens and P. alecto species. In cases where orthologs were not available, we used blast using the peptide sequence as a query to detect the possible source of P. alecto protein. Protein abundances were obtained by summing the abundances of peptides derived from them. We identified differentially expressed proteins by first performing median normalization on all samples from WI-38 and PaLung cells (total six samples), followed by a Student’s t-test for all proteins abundances. p-values were corrected using the false discovery rate method of Benjamini-Hochberg.

For the mitochondrial protein-protein interaction network in Figure 2C, we first obtained a list of all known mitochondrial proteins from the MitoCarta (Calvo et al., 2016) and IMPI databases. These were then input into STRING (Szklarczyk et al., 2019) to identify all high-confidence pairwise protein-protein interactions. We used Cytoscape (Shannon et al., 2003) to both visualize the network and overlay fold change values of detected proteins onto the network. Clustering of the network and Reactome enrichment of identified subnetworks were performed using the ClusterOne app of Cytoscape. For GSEA, the raw abundances of all samples were used as input and GSEA was run against the same GO BP with pathways gene set list, as with RNAseq data.

Metabolomics

For the metabolomics comparison, three independent pairs of WI-38 and PaLung cells were cultured using media conditions as described above in 20% O2 and 5% CO2-humidified atmosphere at 37°C before harvesting. WI-38 and PaLung cells were harvested according to the protocol outlined in the document ACB.1.0.0 provided by Human Metabolome Technologies (HMT Japan). Targeted quantitative analysis was performed by HMT, using capillary electrophoresis mass spectrometry (CE-TOFMS and CE-QqQMS). Absolute abundances (adjusted for cell numbers) were obtained for a total of 116 metabolites (52 and 64 metabolites in the cation and anion modes, respectively). The metabolomics data has been deposited on the Metabolomics Workbench repository (Sud et al., 2016).

Computational flux analysis

Metabolic network reconstruction

A network model of central metabolism in humans was obtained from Mitocore (Smith et al., 2017) (485 reactions, 371 metabolites). This network contains most mitochondrial reactions and pathways, with additional reactions for glycolysis, pentose phosphate pathway, and import and export of amino acids, ions, and other metabolites. The mitocore model also contains a list of genes that catalyze each reaction in a gene-reaction rules table (both mandatory and optional genes). Using this base model, separate context-specific reconstructions for PaLung and WI-38 cell lines were obtained manually using the proteomic and transcriptomic data as follows. The expression levels of all enzymes of the mitocore model were checked in our proteomics and RNAseq datasets, and enzymes were marked as missing if they had <5 counts in two out of three samples (for each PaLung and WI-38) for transcriptomics, and <2000 abundance in two out of three samples for proteomics. Reactions in the mitocore model whose activity depended on the presence of missing genes were iteratively removed from the model while ensuring that the removal of reactions would not result in absence of flux-carrying capacity for ATP production, TCA cycle, or glycolysis. The final reconstructed model for PaLung cells contained 409 reactions and 324 metabolites, and the reconstructed model for WI-38 cells contained 437 reactions and 341 metabolites. We call these PaLung and WI-38 models the ‘reconstructed models’.

Flux sampling

For each species, we first obtained a feasible range of metabolic flux levels for each reaction by performing flux sampling on the ‘reconstructed’ model for that species. To perform flux sampling, we generated 5000 flux vectors through uniform sampling using the COBRA toolbox v3.0 (Heirendt et al., 2019). We call this round of flux sampling the ‘control simulation’ as there are no constraints imposed on the flux of the reactions for Complex I and oxygen consumption, meaning that these reactions are free to assume any flux values that satisfy the steady-state assumption. Next, we repeated the flux sampling process for a case where the PaLung model was forced to have higher flux through the Complex I reaction than the WI-38 model, and lower flux through the oxygen consumption reaction than the WI-38 model. We call this the ‘constrained simulation’. The setting of constraints is explained in the following section.

Constraints for flux sampling

Constraining reaction fluxes in a metabolic network is usually accomplished by setting a lower bound or an upper bound (or both) on the feasible range of the flux values for individual reactions. When a lower bound is set, flux sampling will only output flux vectors where the flux for that reaction is greater than or equal to the set lower bound. Similarly, setting an upper bound ensures that flux sampling will only output flux vectors where the reaction flux is less than or equal to the set upper bound. To ensure that the PaLung model would have higher Complex I activity and lower mitochondrial oxygen consumption than the WI-38 model, we set flux constraints as follows. We first computed the maximum possible fluxes of the Complex I reaction (Reaction ID: CI_MitoCore) in both PaLung and WI-38 models. We then set the lower bound of the PaLung Complex I reaction flux to a value equal to 70% of its theoretical maximum. Similarly, we set the upper bound of the WI-38 Complex I reaction at a value equal to 30% of its theoretical maximum value. This ensured that the PaLung model would have higher flux through the Complex I reaction, in comparison to the WI-38 model. A similar process of identifying maximum possible fluxes and setting lower and upper bounds was followed for the oxygen consumption reaction (Reaction ID: O2tm) to ensure that the lower bound of the reaction in the WI-38 model was higher than the upper bound in the PaLung model. Flux sampling was then performed for both the constrained PaLung model and the constrained WI-38 model to obtain 5000 flux vectors (as with the control simulation). More details about setting flux sampling constraints and the effects of using different constraint values can be found in Appendix 3.

MitoSOX assay and antimycin A treatment

PaLung and WI-38 cells were cultured on six-well plates and treated with either DMSO or antimycin A (1 μΜ) for 16 hr. Antimycin A-treated cells were washed twice with PBS before adding fresh media containing 2.5 μΜ MitoSOX Red. Cells were stained with MitoSOX Red for 60 min and subsequently collected for flow cytometry. Quantification and analysis of the data were performed by FlowJo software.

NADPH/NADP+ measurement assay

Intracellular NADPH/NADP+ were measured using the NADP/NADPH Quantification Kit (ab65349, Abcam) according to the manufacturer’s instructions. Briefly, 6×105 cells were lysed with 350 μl of extraction buffer. For the reaction, 50 μl of the final sample was used. Signal intensities for NADPH were examined by OD measurements at 450 nm using Infinite M200 plate reader (TECAN).

GSH/GSSG measurement assay

Intracellular GSH/GSSG was measured using GSH/GSSG-Glo luminescent assay (Promega) according to the manufacturer’s instructions. Briefly, 2×104 cells in 96-well plates were lysed in the indicated condition.

Mitochondrial OCR measurement

OCR was measured using a Seahorse Bioscience XF96 Extracellular Flux Analyzer (Seahorse Bioscience; RRID: SCR_019545). 2×104 cells were plated into Seahorse tissue culture 96-well plates. Cells were cultured in Seahorse assay media containing 10 mM glucose and 2 mM glutamine and incubated in a CO2-free incubator for an hour before measurement. XF Cell Mito Stress Test Kit was used to analyze mitochondrial metabolic parameters by measuring OCR. Oligomycin (1 μΜ) was injected to determine the oligomycin-independent lack of the OCR. The mitochondrial uncoupler FCCP (1 μΜ) was injected to determine the maximum respiratory capacity. Rotenone (1 μΜ) and antimycin A (1 μΜ) were injected to block Complex I and Complex III of the ETC.

ECAR measurement

ECAR was measured using a Seahorse Bioscience XF96 Extracellular Flux Analyzer (Seahorse Bioscience). 2×104 cells were plated into Seahorse tissue culture 96-well plates. Cells were cultured in Seahorse assay media containing 2 mM glutamine and incubated in a CO2-free incubator for an hour before measurement. XF Cell Glycolysis Stress Test Kit was used to analyze glycolytic metabolic parameters by measuring ECAR. Glucose (10 mM), oligomycin (1 μΜ), and 2-deoxy-D-glucose (50 mM) were injected sequentially.

Acknowledgements

The authors thank Dr. Akshamal M Gamage and Prof. Lena Ho for their insightful commentary on the manuscript. This research is supported by the Singapore Ministry of Education Academic Research Fund Tier 2 grant (MOE2019-T2-1-138) to LTK; Singapore Ministry of Education Academic Research Fund Tier 2 Grant (MOE-T2EP30120-0012) to KI; the Singapore Ministry of Health’s National Medical Research Council grant (NMRC/OFIRG/MOH-000639) to KI; and Duke-NUS Signature Research Programme Block Grants to KI and LTK. Any opinions, findings, conclusions, or recommendations expressed in this material are those of the authors and do not reflect the views of the funding agencies.

Appendix 1

GSEA of proteomics data from mitochondrial fractions without the outlier P1 and W1 samples

Since the samples P1 and W1 were observed to be outliers from the heatmap in Figure 2B of the main text, we repeated the proteomics GSEA, leaving out these two samples and using only P2 and P3 for PaLung and W2 and W3 for WI-38 (n=2). However, removing P1 and W1 did not affect our original results and we still observe that ETC and Complex I in particular are upregulated in the PaLung samples.

Appendix 1—figure 1. Gene set enrichment analysis (GSEA) enrichment plots of the same gene sets as shown in Figure 2D and E of the main text, after removing the outlier proteomic samples P1 and W1 from the analysis (n=2).

Appendix 1—figure 1.

Appendix 1—table 1. Table showing the top 35 gene sets enriched in the PaLung mitochondrial proteomics samples, after removing the outlier samples P1 and W1.

Columns indicate the name of the gene set, size (number of genes in gene set), normalized enrichment score, and the false discovery rate (FDR) value.

NAME SIZE NES FDR q-val
MUSCLE CONTRACTION GOBP GO:0006936 35 –2.301 0.002
NICOTINIC ACETYLCHOLINE RECEPTOR SIGNALING PATHWAY PANTHER PATHWAY P00044 15 –2.191 0.012
REGULATION OF CELL JUNCTION ASSEMBLY GOBP GO:1901888 15 –2.163 0.011
THE CITRIC ACID (TCA) CYCLE AND RESPIRATORY ELECTRON TRANSPORT REACTOME R-HSA-1428517.1 64 –2.154 0.011
HALLMARK_OXIDATIVE_PHOSPHORYLATION MSIGDB_C2 HALLMARK_OXIDATIVE_PHOSPHORYLATION 91 –2.142 0.010
COLLAGEN FORMATION REACTOME DATABASE ID RELEASE 71 1474290 31 –2.096 0.015
RESPIRATORY ELECTRON TRANSPORT, ATP SYNTHESIS BY CHEMIOSMOTIC COUPLING, AND HEAT PRODUCTION BY UNCOUPLING PROTEINS. REACTOME R-HSA-163200.1 32 –2.074 0.017
RESPIRATORY ELECTRON TRANSPORT REACTOME R-HSA-611105.3 32 –2.059 0.019
NADH DEHYDROGENASE COMPLEX ASSEMBLY GOBP GO:0010257 16 –2.059 0.017
EPH-EPHRIN SIGNALING REACTOME DATABASE ID RELEASE 71 2682334 30 –2.057 0.016
COMPLEX I BIOGENESIS REACTOME R-HSA-6799198.1 16 –2.036 0.018
MUSCLE SYSTEM PROCESS GOBP GO:0003012 40 –2.033 0.017
RHO GTPASES ACTIVATE PKNS REACTOME DATABASE ID RELEASE 71 5625740 17 –2.030 0.016
ACTIN FILAMENT-BASED MOVEMENT GOBP GO:0030048 17 –2.022 0.017
ACTOMYOSIN STRUCTURE ORGANIZATION GOBP GO:0031032 21 –2.021 0.016
MITOCHONDRIAL RESPIRATORY CHAIN COMPLEX I ASSEMBLY GOBP GO:0032981 16 –1.999 0.019
INTEGRIN SIGNALLING PATHWAY PANTHER PATHWAY P00034 44 –1.994 0.019
KERATINIZATION GOBP GO:0031424 15 –1.976 0.021
MITOCHONDRIAL ATP SYNTHESIS COUPLED ELECTRON TRANSPORT GOBP GO:0042775 26 –1.971 0.021
MITOCHONDRIAL ELECTRON TRANSPORT, NADH TO UBIQUINONE GOBP GO:0006120 15 –1.968 0.021
MITOCHONDRIAL RESPIRATORY CHAIN COMPLEX ASSEMBLY GOBP GO:0033108 19 –1.966 0.020
CORNIFICATION GOBP GO:0070268 15 –1.963 0.019
SYSTEM PROCESS GOBP GO:0003008 77 –1.953 0.021
MIDBRAIN DEVELOPMENT GOBP GO:0030901 15 –1.953 0.020
COLLAGEN BIOSYNTHESIS AND MODIFYING ENZYMES REACTOME R-HSA-1650814.3 25 –1.949 0.020
KERATINIZATION REACTOME DATABASE ID RELEASE 71 6805567 15 –1.936 0.022
ELECTRON TRANSPORT CHAIN (OXPHOS SYSTEM IN MITOCHONDRIA) WIKIPATHWAYS_20191210 WP111 HOMO SAPIENS 28 –1.933 0.022
NABA_CORE_MATRISOME MSIGDB_C2 NABA_CORE_MATRISOME 34 –1.921 0.025
CELLULAR RESPIRATION GOBP GO:0045333 52 –1.905 0.028
HALLMARK_ESTROGEN_RESPONSE_EARLY MSIGDB_C2 HALLMARK_ESTROGEN_RESPONSE_EARLY 16 –1.902 0.028
EXTRACELLULAR MATRIX ORGANIZATION REACTOME DATABASE ID RELEASE 71 1474244 59 –1.895 0.029
RHO GTPASES ACTIVATE PAKS REACTOME R-HSA-5627123.2 16 –1.863 0.038
INFLAMMATION MEDIATED BY CHEMOKINE AND CYTOKINE SIGNALING PATHWAY PANTHER PATHWAY P00031 30 –1.848 0.043
ELECTRON TRANSPORT CHAIN GOBP GO:0022900 31 –1.842 0.044
ATP SYNTHESIS COUPLED ELECTRON TRANSPORT GOBP GO:0042773 27 –1.832 0.048

Appendix 2

GSEA of mitochondrial fractions using mitochondrial gene sets

Our proteomics data was obtained from mitochondrial fraction samples, and the GSEA in Figure 2 of the main text was originally performed using the entire GO BP gene set list. As an additional measure of redundancy, we repeated our GSEA for the mitochondrial proteomics samples using a gene set list specifically curated for mitochondrial fraction analysis. The gene sets for this analysis were obtained from the MitoCarta 3.0 database (gene set file: MitoPathways3.0.gmx). The results from this analysis also agree with our findings from using the entire GO BP gene set list, that OxPhos and specifically Complex I proteins are upregulated in PaLung mitochondrial samples.

Appendix 2—figure 1. Gene set enrichment analysis (GSEA) enrichment plots of the gene sets that are biological equivalents of the gene sets shown in Figure 2D and E of the main text.

Appendix 2—figure 1.

This analysis was performed using the MitoCarta 3.0 gene set list instead of the Gene Ontology Biological Process (GO BP) gene set list used to generate Figure 2D and E.

Appendix 2—table 1. Table showing the top gene sets enriched in the PaLung mitochondrial proteomics samples, when gene set enrichment analysis (GSEA) was performed using the MitoCarta 3.0 gene set list instead of the Gene Ontology Biological Process (GO BP) gene set.

Columns indicate the name of the gene set, size (number of genes in gene set), normalized enrichment score, and the false discovery rate (FDR) value.

NAME SIZE NES FDR q-val
OXPHOS_SUBUNITS 34 –2.186 0.002
OXPHOS 41 –2.166 0.001
CARBOHYDRATE_METABOLISM 36 –2.004 0.005
TRANSLATION 16 –1.987 0.005
COMPLEX_I 17 –1.966 0.005
CI_SUBUNITS 15 –1.892 0.007
FATTY_ACID_OXIDATION 20 –1.864 0.007
METALS_AND_COFACTORS 30 –1.794 0.011
METABOLISM 153 –1.782 0.011
AMINO_ACID_METABOLISM 33 –1.763 0.011
MITOCHONDRIAL_CENTRAL_DOGMA 24 –1.609 0.030
TCA_CYCLE 15 –1.510 0.053
LIPID_METABOLISM 43 –1.480 0.057
PROTEIN_IMPORT_SORTING_AND_HOMEOSTASIS 24 –1.200 0.223
PROTEIN_HOMEOSTASIS 18 –1.053 0.377

Appendix 3

Setting constraints for flux sampling

Our primary goal with the flux simulations is to establish constraints that explore the potential metabolic implications of our key observations from omics and metabolic measurements: bats have upregulated Complex I genes but lower mitochondrial oxygen consumption. Toward this goal, we first set constraints on the Complex I reaction (CI_MitoCore) and the mitochondrial oxygen transport reaction (O2tm) in our metabolic models of PaLung (P model) and WI-38 (W model). In an ideal scenario, these two reactions are completely independent of each other in flux space and setting constraints/thresholds on one reaction does not affect the feasible flux space of the other reaction. However, in our case we observed that constraining one reaction also limits the feasible flux space of the other reaction. Hence, we follow the below protocol in setting our constraints, to ensure that the Complex I flux of bats is higher than humans and vice versa for the O2 flux. We call this the 30-70 protocol as the bounds are set to 30% or 70% of the feasible flux range for each reaction.

  1. Compute the minimum and maximum flux possible through CI_MitoCore for the two models. These values are designated as [p_c1_min, p_c1_max] (for PaLung) and [w_c1_min, w_c1_max] (for WI-38).

  2. Constrain the CI reaction in the P model to have a lower bound of p_c1_min + 0.7*(p_c1_max- p_c1_min).

  3. Constrain the CI reaction in the W model to have an upper bound of w_c1_min + 0.3*(w_c1_max- w_c1_min).

  4. Now that the CI reactions have been constrained, compute the new minimum and maximum flux possible through O2 for the two models. These values are designated as [p_o2_min, p_o2_max] (for PaLung) and [w_o2_min, w_o2_max] (for WI-38).

  5. Constrain the O2 reaction in the P model to have an upper bound of p_o2_min + 0.3*(p_o2_max- p_o2_min). Designate this upper bound as p_o2_ub.

  6. To avoid overlap of flux ranges and to ensure that the lower bound of the W model O2 reaction is greater than the upper bound of the P model O2 reaction, constrain the O2 reaction in the W model to have a lower bound equal to the higher value between the following two values.

    1. w_o2_min +0.7*(w_o2_max- w_o2_min).

    2. p_o2_ub.

Our methodology may raise the question of what happens if we were to follow the procedure in reverse and first constrain the O2 reaction and then the CI reaction following the same protocol. To answer this question, we performed this simulation as well. As seen in the figure below, this also results in the Complex II reaction having low-to-negative values in the bat model (not seen with the human model).

Appendix 3—figure 1. Flux sampling histograms of the P and W metabolic models in the unconstrained control (left column) and the constrained (right column) cases.

Appendix 3—figure 1.

The P fluxes are in blue and the W fluxes in red. The first two rows show the constrained reactions (Complex I and O2), while the third row shows the flux histograms of the Complex II reaction.

In addition, we also provide Appendix 3—table 1. Appendix 3—table 1 shows the minimum and maximum flux possible through the Complex I and O2 reactions when different constraints/combinations of constraints are imposed on the P and W models.

Appendix 3—table 1. Minimum and maximum flux values possible for the Complex I and mitochondrial O2 transport reaction when different constraints are applied to the P and W metabolic models.

Constraint description (all follow the 30-70 protocols) Complex 1 Oxygen transport
Bat (P model) Human (W model) Bat (P model) Human (W model)
Min Max Min Max Min Max Min Max
Control simulation – no constraints 0 41.43 0 41.44 0 19.8 0 19.8
Ideal target flux range expected with the 30-70 protocol, when CI and O2 are independent of each other 29 41.43 0 12.43 0 5.94 13.86 19.8
Constraining only the Complex I reaction 29 41.43 0 12.43 13.58 19.8 0 19.8
Constraining only the O2 reaction 0 13.74 0 41.43 0 5.94 13.86 19.8
Constraining CI first, then constraining O2, without avoiding flux range overlap 29 32.72 0 12.43 13.58 15.45 13.86 19.8
Constraining CI first, then constraining O2, avoiding flux range overlap 29 32.72 0 12.43 13.58 15.45 15.45 19.8
Constraining O2 first, then constraining CI, without avoiding flux range overlap 9.61 13.73 0 12.43 3.88 5.94 13.86 19.8
Constraining O2 first, then constraining CI, avoiding flux range overlap 9.61 13.73 0 9.61 3.88 5.94 13.86 19.8

We also performed further flux sampling simulations to explore the effect of using the 30-70 protocol for threshold percentages, and whether the results would be robust to other threshold percentages. We repeated our original simulations following our original protocol (constrain C1 first and then O2), with thresholds of 20-80, 40-60, and 50-50 Appendix 3—figure 2 . In all cases, we observed that the Complex II reaction has low-to-negative values in the bat model compared to the human model to different degrees. We thus conclude that this observation is robust to the choice of constraint thresholds.

Our flux sampling scripts can be found at https://github.com/narendrasuhas/PalungWI38FluxSim, (copy archived at Narendrasuhas, 2024). The script contains a variable called minFrac, which can be set to values between 0 and 1 for setting different thresholds. For example, a minFrac value of 0.3 ensures a 30-70 protocol, while a minFrac value of 0.2 ensures a 20-80 protocol.

Appendix 3—figure 2. Flux sampling histograms of the P and W metabolic models under different threshold values for constraints.

Appendix 3—figure 2.

Each row corresponds to a constraint scheme (unconstrained, 20-80, 30-70, 40-60, and 50-50), and was obtained by running our script with different minFrac values. Within each panel, the P model fluxes are in blue and the W model fluxes in red. The leftmost column shows the flux histograms of the Complex I reaction, the middle column shows flux histograms of the mitochondrial O2 transport reaction, and the rightmost column shows the flux histograms of the Complex II reaction.

Funding Statement

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

Contributor Information

Yoko Itahana, Email: yoko.itahana@duke-nus.edu.sg.

Koji Itahana, Email: koji.itahana@duke-nus.edu.sg.

Lisa Tucker-Kellogg, Email: tuckerNUS@gmail.com.

Pankaj Kapahi, Buck Institute for Research on Aging, United States.

Pankaj Kapahi, Buck Institute for Research on Aging, United States.

Funding Information

This paper was supported by the following grants:

  • Ministry of Education - Singapore MOE2019-T2-1-138 to Lisa Tucker-Kellogg.

  • Ministry of Education - Singapore MOE-T2EP30120-0012 to Koji Itahana.

  • National Medical Research Council NMRC/OFIRG/MOH-000639 to Koji Itahana.

  • Duke-NUS Medical School Block Grants to Koji Itahana, Lisa Tucker-Kellogg.

Additional information

Competing interests

No competing interests declared.

Author contributions

Conceptualization, Formal analysis, Visualization, Writing – original draft.

Conceptualization, Investigation, Writing – original draft.

Investigation, Visualization, Writing – original draft.

Investigation.

Formal analysis, Investigation.

Conceptualization.

Conceptualization, Investigation, Writing – original draft.

Conceptualization, Writing – original draft.

Conceptualization, Formal analysis, Writing – original draft.

Additional files

Supplementary file 1. Genes that pass our differential expression cutoffs (false discovery rate [FDR] < 0.05; |log fold change| > 1) in PaLung vs WI-38 samples from whole-cell transcriptomics data.

Differential expression analysis was performed using the DESeq2 pipeline. Fold changes are indicated as PaLung/WI-38.

elife-94007-supp1.xlsx (657.9KB, xlsx)
Supplementary file 2. Biological pathways upregulated in PaLung cells from gene set enrichment analysis (GSEA) of transcriptomics data.

GSEA was performed on the transcriptomics data using (PI) value as a metric. The table below lists the pathways detected as upregulated in PaLung cells (compared to WI-38 cells) and associated enrichment metrics.

elife-94007-supp2.xlsx (245.8KB, xlsx)
Supplementary file 3. Biological pathways upregulated in WI-38 cells from gene set enrichment analysis (GSEA) of transcriptomics data.

GSEA was performed on the transcriptomics data using (PI) value as a metric. The table below lists the pathways detected as upregulated in WI-38 cells (compared to PaLung cells) and associated enrichment metrics.

elife-94007-supp3.xlsx (158.8KB, xlsx)
Supplementary file 4. Differentially expressed (DE) mitochondrial proteins in PaLung vs WI-38 samples from mitochondrial proteomics data.

405 DE proteins were first identified using a Student’s t-test on median-corrected protein abundances from the mitochondrial samples of PaLung and WI-38. Of the 405 DE proteins, 127 were identified to be core mitochondrial proteins (as defined by MitoCarta and IMPI datasets) and are listed in this sheet. Fold changes are indicated as PaLung/WI-38.

elife-94007-supp4.xlsx (14.2KB, xlsx)
Supplementary file 5. Biological pathways upregulated in PaLung cells from gene set enrichment analysis (GSEA) of proteomics data.

GSEA was performed on the proteomics data using protein abundances as input. The table below lists the pathways detected as upregulated in PaLung cells (compared to WI-38 cells) and associated enrichment metrics.

elife-94007-supp5.xlsx (56.1KB, xlsx)
Supplementary file 6. Biological pathways upregulated in WI-38 cells from gene set enrichment analysis (GSEA) of proteomics data.

GSEA was performed on the proteomics data using protein abundances as input. The table below lists the pathways detected as upregulated in WI-38 cells (compared to PaLung cells) and associated enrichment metrics.

elife-94007-supp6.xlsx (81KB, xlsx)
Supplementary file 7. Metabolic model for PaLung cells.

A metabolic flux model was constructed for the central carbon metabolism of PaLung cells by overlaying proteomic and transcriptomic information onto the existing mitocore model from literature.

elife-94007-supp7.xlsx (57.3KB, xlsx)
Supplementary file 8. Metabolic model for WI-38 cells.

A metabolic flux model was constructed for the central carbon metabolism of WI-38 cells by overlaying proteomic and transcriptomic information onto the existing mitocore model from literature.

elife-94007-supp8.xlsx (60.4KB, xlsx)
Supplementary file 9. Flux sampling results comparing flux distributions in the constrained PaLung and WI-38 models.

Flux sampling was performed with 5000 flux vectors for the PaLung and WI-38 metabolic models each. The flux histograms for each reaction were compared across the two models and the following statistics were extracted from the histograms.

elife-94007-supp9.xlsx (67.8KB, xlsx)
Supplementary file 10. Absolute metabolite quantification in PaLung and WI-38 cells.

Absolute concentrations of metabolites detected in PaLung and WI-38 cells by Human Metabolome Technologies (HMT).

elife-94007-supp10.xlsx (40.7KB, xlsx)
MDAR checklist

Data availability

Transcriptomic data are deposited in the NCBI GEO database (GSE215934). Proteomic data are deposited in the ProteomeXchange database (PXD043121) and in the jPOST repository (JPST001821). Metabolomics data are uploaded to the Metabolomics Workbench database (ST002743). The Matlab script used for flux sampling can be found at here (copy archived at Jagannathan, 2023). Other data supporting the findings of this study are available within the article and its supplementary materials.

The following datasets were generated:

Koh J, Irving A, Itahana Y, Lee Y, Itahana K, Wang L, Suhas Jagannathan N, Tucker-Kellogg L. 2024. RNAseq comparison of lung fibroblasts from Pteropus alecto (PaLung cell line) and Homo sapiens (WI-38 cell line) NCBI Gene Expression Omnibus. GSE215934

Itahana K. 2024. Multi-omic analysis of bat versus human fibroblasts reveals altered central metabolism. jPOST repository. JPST001821

Koh J, Jagannathan MS, Sobota RM, Tucker-Kellogg L, Itahana Y, Itahana K. 2024. Comparison of human and bat metabolism. ProteomeXchange. PXD043121

Koh J, Jagannathan NS, Tucker-Kellogg L, Itahana Y, Itahana K. 2024. Metabolomics comparison of lung fibroblasts from Pteropus alecto and Homo sapiens. Metabolomics Workbench. ST002743

References

  1. Ahn M, Anderson DE, Zhang Q, Tan CW, Lim BL, Luko K, Wen M, Chia WN, Mani S, Wang LC, Ng JHJ, Sobota RM, Dutertre C-A, Ginhoux F, Shi Z-L, Irving AT, Wang L-F. Dampened NLRP3-mediated inflammation in bats and implications for a special viral reservoir host. Nature Microbiology. 2019;4:789–799. doi: 10.1038/s41564-019-0371-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Atkinson DE, Walton GM. Adenosine triphosphate conservation in metabolic regulation. Rat liver citrate cleavage enzyme. The Journal of Biological Chemistry. 1967;242:3239–3241. [PubMed] [Google Scholar]
  3. Aurich MK, Fleming RMT, Thiele I. A systems approach reveals distinct metabolic strategies among the NCI-60 cancer cell lines. PLOS Computational Biology. 2017;13:e1005698. doi: 10.1371/journal.pcbi.1005698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Austad SN, Fischer KE. Mammalian aging, metabolism, and ecology: evidence from the bats and marsupials. Journal of Gerontology. 1991;46:B47–B53. doi: 10.1093/geronj/46.2.b47. [DOI] [PubMed] [Google Scholar]
  5. Austad SN. Methusaleh’s Zoo: how nature provides us with clues for extending human health span. Journal of Comparative Pathology. 2010;142 Suppl 1:S10–S21. doi: 10.1016/j.jcpa.2009.10.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bisbach CM, Hass DT, Robbings BM, Rountree AM, Sadilek M, Sweet IR, Hurley JB. Succinate can shuttle reducing power from the hypoxic retina to the O2-rich pigment epithelium. Cell Reports. 2020;31:107606. doi: 10.1016/j.celrep.2020.107606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bordbar A, Mo ML, Nakayasu ES, Schrimpe-Rutledge AC, Kim Y-M, Metz TO, Jones MB, Frank BC, Smith RD, Peterson SN, Hyduke DR, Adkins JN, Palsson BO. Model-driven multi-omic data analysis elucidates metabolic immunomodulators of macrophage activation. Molecular Systems Biology. 2012;8:558. doi: 10.1038/msb.2012.21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Brown JCL, McClelland GB, Faure PA, Klaiman JM, Staples JF. Examining the mechanisms responsible for lower ROS release rates in liver mitochondria from the long-lived house sparrow (Passer domesticus) and big brown bat (Eptesicus fuscus) compared to the short-lived mouse (Mus musculus) Mechanisms of Ageing and Development. 2009;130:467–476. doi: 10.1016/j.mad.2009.05.002. [DOI] [PubMed] [Google Scholar]
  9. Brunet-Rossinni AK. Reduced free-radical production and extreme longevity in the little brown bat (Myotis lucifugus) versus two non-flying mammals. Mechanisms of Ageing and Development. 2004;125:11–20. doi: 10.1016/j.mad.2003.09.003. [DOI] [PubMed] [Google Scholar]
  10. Brunet-Rossinni AK, Austad SN. Ageing studies on bats: a review. Biogerontology. 2004;5:211–222. doi: 10.1023/B:BGEN.0000038022.65024.d8. [DOI] [PubMed] [Google Scholar]
  11. Buffenstein R, Edrey YH, Yang T, Mele J. The oxidative stress theory of aging: embattled or invincible? Insights from non-traditional model organisms. Age. 2008;30:99–109. doi: 10.1007/s11357-008-9058-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Calvo SE, Clauser KR, Mootha VK. MitoCarta2.0: an updated inventory of mammalian mitochondrial proteins. Nucleic Acids Research. 2016;44:D1251–D1257. doi: 10.1093/nar/gkv1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Chen Y, Fan H, Wang S, Tang G, Zhai C, Shen L. Ferroptosis: a novel therapeutic target for ischemia-reperfusion injury. Frontiers in Cell and Developmental Biology. 2021;9:688605. doi: 10.3389/fcell.2021.688605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Chini CCS, Tarragó MG, Chini EN. NAD and the aging process: Role in life, death and everything in between. Molecular and Cellular Endocrinology. 2017;455:62–74. doi: 10.1016/j.mce.2016.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Chionh YT, Cui J, Koh J, Mendenhall IH, Ng JHJ, Low D, Itahana K, Irving AT, Wang LF. High basal heat-shock protein expression in bats confers resistance to cellular heat/oxidative stress. Cell Stress & Chaperones. 2019;24:835–849. doi: 10.1007/s12192-019-01013-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Chouchani ET, Pell VR, Gaude E, Aksentijević D, Sundier SY, Robb EL, Logan A, Nadtochiy SM, Ord ENJ, Smith AC, Eyassu F, Shirley R, Hu C-H, Dare AJ, James AM, Rogatti S, Hartley RC, Eaton S, Costa ASH, Brookes PS, Davidson SM, Duchen MR, Saeb-Parsy K, Shattock MJ, Robinson AJ, Work LM, Frezza C, Krieg T, Murphy MP. Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature. 2014;515:431–435. doi: 10.1038/nature13909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Clément MV, Pervaiz S. Intracellular superoxide and hydrogen peroxide concentrations: a critical balance that determines survival or death. Redox Report. 2001;6:211–214. doi: 10.1179/135100001101536346. [DOI] [PubMed] [Google Scholar]
  18. Crameri G, Todd S, Grimley S, McEachern JA, Marsh GA, Smith C, Tachedjian M, De Jong C, Virtue ER, Yu M, Bulach D, Liu J-P, Michalski WP, Middleton D, Field HE, Wang L-F. Establishment, immortalisation and characterisation of pteropid bat cell lines. PLOS ONE. 2009;4:e8266. doi: 10.1371/journal.pone.0008266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Di Meo S, Reed TT, Venditti P, Victor VM. Harmful and beneficial role of ROS. Oxidative Medicine and Cellular Longevity. 2016;2016:1–3. doi: 10.1155/2016/7909186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Ezzamouri B, Rosario D, Bidkhori G, Lee S, Uhlen M, Shoaie S. Metabolic modelling of the human gut microbiome in type 2 diabetes patients in response to metformin treatment. NPJ Systems Biology and Applications. 2023;9:2. doi: 10.1038/s41540-022-00261-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Foley NM, Hughes GM, Huang Z, Clarke M, Jebb D, Whelan CV, Petit EJ, Touzalin F, Farcy O, Jones G, Ransome RD, Kacprzyk J, O’Connell MJ, Kerth G, Rebelo H, Rodrigues L, Puechmaille SJ, Teeling EC. Growing old, yet staying young: The role of telomeres in bats’ exceptional longevity. Science Advances. 2018;4:eaao0926. doi: 10.1126/sciadv.aao0926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Heirendt L, Arreckx S, Pfau T, Mendoza SN, Richelle A, Heinken A, Haraldsdóttir HS, Wachowiak J, Keating SM, Vlasov V, Magnusdóttir S, Ng CY, Preciat G, Žagare A, Chan SHJ, Aurich MK, Clancy CM, Modamio J, Sauls JT, Noronha A, Bordbar A, Cousins B, El Assal DC, Valcarcel LV, Apaolaza I, Ghaderi S, Ahookhosh M, Ben Guebila M, Kostromins A, Sompairac N, Le HM, Ma D, Sun Y, Wang L, Yurkovich JT, Oliveira MAP, Vuong PT, El Assal LP, Kuperstein I, Zinovyev A, Hinton HS, Bryant WA, Aragón Artacho FJ, Planes FJ, Stalidzans E, Maass A, Vempala S, Hucka M, Saunders MA, Maranas CD, Lewis NE, Sauter T, Palsson BØ, Thiele I, Fleming RMT. Creation and analysis of biochemical constraint-based models using the COBRA Toolbox v.3.0. Nature Protocols. 2019;14:639–702. doi: 10.1038/s41596-018-0098-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Huang Z, Jebb D, Teeling EC. Blood miRNomes and transcriptomes reveal novel longevity mechanisms in the long-lived bat, Myotis myotis. BMC Genomics. 2016;17:906. doi: 10.1186/s12864-016-3227-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Huang Z, Whelan CV, Foley NM, Jebb D, Touzalin F, Petit EJ, Puechmaille SJ, Teeling EC. Longitudinal comparative transcriptomics reveals unique mechanisms underlying extended healthspan in bats. Nature Ecology & Evolution. 2019;3:1110–1120. doi: 10.1038/s41559-019-0913-3. [DOI] [PubMed] [Google Scholar]
  25. Jagannathan NJ. PalungWI38FluxSim. swh:1:rev:ef9950f3820fd0612f5f1403facc8a7d262f0ba0Software Heritage. 2023 https://archive.softwareheritage.org/swh:1:dir:176e93fe3b5e1e4535fe97cb30bc438891f7a5fd;origin=https://github.com/narendrasuhas/PalungWI38FluxSim;visit=swh:1:snp:5f542abbf376c3afefd65f8336aabc09afc3904f;anchor=swh:1:rev:ef9950f3820fd0612f5f1403facc8a7d262f0ba0
  26. Jansma J, El Aidy S. Understanding the host-microbe interactions using metabolic modeling. Microbiome. 2021;9:16. doi: 10.1186/s40168-020-00955-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Kacprzyk J, Hughes GM, Palsson-McDermott EM, Quinn SR, Puechmaille SJ, O’Neill LAJ, Teeling EC. A potent anti-inflammatory response in bat macrophages may be linked to extended longevity and viral tolerance. Acta Chiropterologica. 2017;19:219–228. doi: 10.3161/15081109ACC2017.19.2.001. [DOI] [Google Scholar]
  28. Kelm DH, Simon R, Kuhlow D, Voigt CC, Ristow M. High activity enables life on a high-sugar diet: blood glucose regulation in nectar-feeding bats. Proceedings. Biological Sciences. 2011;278:3490–3496. doi: 10.1098/rspb.2011.0465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Koh J, Itahana Y, Mendenhall IH, Low D, Soh EXY, Guo AK, Chionh YT, Wang LF, Itahana K. ABCB1 protects bat cells from DNA damage induced by genotoxic compounds. Nature Communications. 2019;10:2820. doi: 10.1038/s41467-019-10495-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Liao S, Apaijai N, Luo Y, Wu J, Chunchai T, Singhanat K, Arunsak B, Benjanuwattra J, Chattipakorn N, Chattipakorn SC. Cell death inhibitors protect against brain damage caused by cardiac ischemia/reperfusion injury. Cell Death Discovery. 2021;7:312. doi: 10.1038/s41420-021-00698-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biology. 2014;15:550. doi: 10.1186/s13059-014-0550-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Maina JN. What it takes to fly: the structural and functional respiratory refinements in birds and bats. The Journal of Experimental Biology. 2000;203:3045–3064. doi: 10.1242/jeb.203.20.3045. [DOI] [PubMed] [Google Scholar]
  33. Mittler R. ROS Are Good. Trends in Plant Science. 2017;22:11–19. doi: 10.1016/j.tplants.2016.08.002. [DOI] [PubMed] [Google Scholar]
  34. Narendrasuhas Palungwi38Fluxsim. swh:1:rev:89a79623290fe89ebaa7bb42eb1dd20e782a4c6aSoftware Heritage. 2024 https://archive.softwareheritage.org/swh:1:dir:96cb562ec4c8fd415cfa533aace904c48e4d7e29;origin=https://github.com/narendrasuhas/PalungWI38FluxSim;visit=swh:1:snp:2b10ba39d7a8a4dfd2618a3427509616fd19ac98;anchor=swh:1:rev:89a79623290fe89ebaa7bb42eb1dd20e782a4c6a
  35. O’Brien KP, Remm M, Sonnhammer ELL. Inparanoid: a comprehensive database of eukaryotic orthologs. Nucleic Acids Research. 2005;33:D476–D480. doi: 10.1093/nar/gki107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Patil KR, Akesson M, Nielsen J. Use of genome-scale microbial models for metabolic engineering. Current Opinion in Biotechnology. 2004;15:64–69. doi: 10.1016/j.copbio.2003.11.003. [DOI] [PubMed] [Google Scholar]
  37. Podlutsky AJ, Khritankov AM, Ovodov ND, Austad SN. A new field record for bat longevity. The Journals of Gerontology. Series A, Biological Sciences and Medical Sciences. 2005;60:1366–1368. doi: 10.1093/gerona/60.11.1366. [DOI] [PubMed] [Google Scholar]
  38. Ravi S, Gunawan R. ΔFBA-Predicting metabolic flux alterations using genome-scale metabolic models and differential transcriptomic data. PLOS Computational Biology. 2021;17:e1009589. doi: 10.1371/journal.pcbi.1009589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Robinson MD, McCarthy DJ, Smyth GK. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26:139–140. doi: 10.1093/bioinformatics/btp616. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Seluanov A, Gladyshev VN, Vijg J, Gorbunova V. Mechanisms of cancer resistance in long-lived mammals. Nature Reviews. Cancer. 2018;18:433–441. doi: 10.1038/s41568-018-0004-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Shannon P, Markiel A, Ozier O, Baliga NS, Wang JT, Ramage D, Amin N, Schwikowski B, Ideker T. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Research. 2003;13:2498–2504. doi: 10.1101/gr.1239303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Shields HJ, Traa A, Van Raamsdonk JM. Beneficial and detrimental effects of reactive oxygen species on lifespan: a comprehensive review of comparative and experimental studies. Frontiers in Cell and Developmental Biology. 2021;9:628157. doi: 10.3389/fcell.2021.628157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Smid M, Coebergh van den Braak RRJ, van de Werken HJG, van Riet J, van Galen A, de Weerd V, van der Vlugt-Daane M, Bril SI, Lalmahomed ZS, Kloosterman WP, Wilting SM, Foekens JA, IJzermans JNM, MATCH study group. Martens JWM, Sieuwerts AM. Gene length corrected trimmed mean of M-values (GeTMM) processing of RNA-seq data performs similarly in intersample analyses while improving intrasample comparisons. BMC Bioinformatics. 2018;19:236. doi: 10.1186/s12859-018-2246-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Smith AC, Eyassu F, Mazat JP, Robinson AJ. MitoCore: a curated constraint-based model for simulating human central metabolism. BMC Systems Biology. 2017;11:114. doi: 10.1186/s12918-017-0500-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Subramanian A, Tamayo P, Mootha VK, Mukherjee S, Ebert BL, Gillette MA, Paulovich A, Pomeroy SL, Golub TR, Lander ES, Mesirov JP. Gene set enrichment analysis: A knowledge-based approach for interpreting genome-wide expression profiles. PNAS. 2005;102:15545–15550. doi: 10.1073/pnas.0506580102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Sud M, Fahy E, Cotter D, Azam K, Vadivelu I, Burant C, Edison A, Fiehn O, Higashi R, Nair KS, Sumner S, Subramaniam S. Metabolomics workbench: An international repository for metabolomics data and metadata, metabolite standards, protocols, tutorials and training, and analysis tools. Nucleic Acids Research. 2016;44:D463–D470. doi: 10.1093/nar/gkv1042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Szklarczyk D, Gable AL, Lyon D, Junge A, Wyder S, Huerta-Cepas J, Simonovic M, Doncheva NT, Morris JH, Bork P, Jensen LJ, Mering C. STRING v11: protein-protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Research. 2019;47:D607–D613. doi: 10.1093/nar/gky1131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Thomas SP, Suthers RA. The physiology and energetics of bat flight. Journal of Experimental Biology. 1972;57:317–335. doi: 10.1242/jeb.57.2.317. [DOI] [Google Scholar]
  49. Thomas SP. Metabolism during flight in two species of bats, Phyllostomus hastatus and Pteropus gouldii. The Journal of Experimental Biology. 1975;63:273–293. doi: 10.1242/jeb.63.1.273. [DOI] [PubMed] [Google Scholar]
  50. Ungvari Z, Buffenstein R, Austad SN, Podlutsky A, Kaley G, Csiszar A. Oxidative stress in vascular senescence: lessons from successfully aging species. Frontiers in Bioscience. 2008;13:5056–5070. doi: 10.2741/3064. [DOI] [PubMed] [Google Scholar]
  51. Ursini F, Maiorino M. Lipid peroxidation and ferroptosis: The role of GSH and GPx4. Free Radical Biology & Medicine. 2020;152:175–185. doi: 10.1016/j.freeradbiomed.2020.02.027. [DOI] [PubMed] [Google Scholar]
  52. Voigt CC, Speakman JR. Nectar‐feeding bats fuel their high metabolism directly with exogenous carbohydrates. Functional Ecology. 2007;21:913–921. doi: 10.1111/j.1365-2435.2007.01321.x. [DOI] [Google Scholar]
  53. Vyssokikh MY, Holtze S, Averina OA, Lyamzaev KG, Panteleeva AA, Marey MV, Zinovkin RA, Severin FF, Skulachev MV, Fasel N, Hildebrandt TB, Skulachev VP. Mild depolarization of the inner mitochondrial membrane is a crucial component of an anti-aging program. PNAS. 2020;117:6491–6501. doi: 10.1073/pnas.1916414117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Wang LF, Walker PJ, Poon LLM. Mass extinctions, biodiversity and mitochondrial function: are bats “special” as reservoirs for emerging viruses? Current Opinion in Virology. 2011;1:649–657. doi: 10.1016/j.coviro.2011.10.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Wilkinson GS, South JM. Life history, ecology and longevity in bats. Aging Cell. 2002;1:124–131. doi: 10.1046/j.1474-9728.2002.00020.x. [DOI] [PubMed] [Google Scholar]
  56. Xiao Y, Hsiao TH, Suresh U, Chen HIH, Wu X, Wolf SE, Chen Y. A novel significance score for gene selection and ranking. Bioinformatics. 2014;30:801–807. doi: 10.1093/bioinformatics/btr671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Xie B-S, Wang Y-Q, Lin Y, Mao Q, Feng J-F, Gao G-Y, Jiang J-Y. Inhibition of ferroptosis attenuates tissue damage and improves long-term outcomes after traumatic brain injury in mice. CNS Neuroscience & Therapeutics. 2019;25:465–475. doi: 10.1111/cns.13069. [DOI] [PMC free article] [PubMed] [Google Scholar]

Editor's evaluation

Pankaj Kapahi 1

This study analyzed the metabolism of bat cells versus human cells through a comprehensive multi-omics approach, focusing on the black flying fox fruit bat. Findings revealed that bat cells have higher expression levels of Complex I in the electron transport chain but a lower oxygen consumption rate, suggesting a unique metabolic state similar to ischemia. Despite higher levels of mitochondrial reactive oxygen species, bat cells displayed greater antioxidant reserves and resilience to metabolic stress, including glucose deprivation and ferroptosis, highlighting fundamental metabolic differences supporting bats' increased longevity and disease resistance. The study is compelling and provides solid evidence to back the hypothesis.

Decision letter

Editor: Pankaj Kapahi1

In the interests of transparency, eLife publishes the most substantive revision requests and the accompanying author responses.

[Editors' note: this paper was reviewed by Review Commons.]

eLife. 2024 Jul 22;13:e94007. doi: 10.7554/eLife.94007.sa2

Author response


1. General Statements [optional]

We thank all three reviewers for their valuable comments and sincerely appreciate the diligence and their attention to detail in reading our manuscript “Multi-omic analysis of bat versus human fibroblasts reveals altered central metabolism.” All three reviewers agreed that our work is significant, and reviewers described the strengths of our work as follows: novel insight into bat metabolism, providing a first-of-its-kind comprehensive study using multi-omics approaches and modelling for inter-species comparison, and providing a valuable stepping stone for further research on bat metabolism.

The reviewers also raised individual questions about specific points in our work and made suggestions for improving the strength of our claims. We agree with the reviewers' assessments in most cases and will incorporate their suggestions into the text, analyses, and interpretation of our work. Our point-by-point responses explain how we will address each of the reviewers’ concerns, on both the wet bench and computational aspects of our work. These include redoing the following: glucose deprivation cell death assay, transcriptomics analysis using current bioinformatics pipelines, GSEA analysis for both transcriptomics and proteomics data, and flux sampling using modified constraints. Several requests for additional information have been addressed by providing the relevant rationale/justification in our responses. Reviewer 2 suggested biochemical validation of the upregulation of Complex I in bat cells. However, it remains challenging to perform such validation using qPCR or Western blotting in the bat species for reasons that are detailed in our responses. Therefore, following Reviewer 1's suggestion, we have decided to moderate our claims in this manuscript.

2. Description of the planned revisions

Reviewer 1 Major comments

1. Regarding Figure 1A, the authors mention 'n = 3' for a single cell line. Does this refer to three different passages or three independent experiments? Please provide a more detailed description to clarify.

We would like to thank the reviewer for the comment. We apologize for not providing sufficient descriptions of the cells and experiments. Our datasets in Figure 1 were generated from three independent experiments, comparing the cell lines WI-38 (W) and PaLung (P) respectively: W1 vs P1, W2 vs P2, and W3 vs P3. These sets of cells were harvested at different passages in each experiment. Cells were generally subcultured at a 1:2 ratio and they were collected at passage numbers 26 (W1), 30 (W2), and 35 (W3) or 5 (P1), 6 (P2) and 7 (P3). We will provide these details in the Material and Methods section.

2. In relation to Figures 1C and 1D, the authors state in the figure legend that the 'GSEA analysis identifies Respiratory electron transport and Cellular response to hypoxia as the top metabolic pathways that are differentially regulated between PaLung and WI-38 cells.' (Lines 140-144). However, the criteria for selecting these terms as the top metabolic pathways is not clear. In the lists in Supplementary Tables 2 and 3, the authors' proposed term, 'Respiratory electron transport,' is ranked 126th, and 'Cellular response to hypoxia' is ranked 79th. Conversely, terms related to the TCA cycle are ranked 66th and 82nd, and another term that seems to be related to hypoxia, 'OXYGEN-DEPENDENT PROLINE HYDROXYLATION OF HYPOXIA-INDUCIBLE FACTOR Α,' is ranked 62nd. Could the authors please provide a clarification for their choice of 'Respiratory electron transport' and 'Cellular response to hypoxia' as the top metabolic pathways?

We thank the reviewer for raising this question. When presenting results in our original submission, we chose to display gene sets that satisfied the following requirements –

1) The gene set pertains to central metabolism (defined as being a sub-gene set of the superfamily Cellular Metabolic Process GO:0044237)

2) The gene set satisfies the FDR < 0.25 threshold commonly used for GSEA analysis (as opposed to p < 0.05 used in conventional statistics). Justification for why a higher FDR threshold is used for GSEA can be found on the GSEA FAQ website.

(https://software.broadinstitute.org/cancer/software/gsea/wiki/index.php/FAQ)

3) Has a nominal enrichment score with absolute value >= 1

4) When multiple similar gene sets (by biological pathway) satisfy the above constraints, we pick the largest gene set (by gene count) to get a more high-level comparison of metabolism

It was from these criteria that we originally displayed the respiratory electron transport gene set (over the similar TCA cycle gene sets, although the TCA gene sets were ranked higher), and the cellular response to hypoxia gene set. In response to this reviewer’s later comments about the bioinformatics pipeline used (minor comment 1), we propose to redo our RNAseq analysis using updated methods such as STAR. We will also redo our GSEA using the newly updated results from the improved pipeline and will follow the same steps listed above to identify the core metabolic processes that are altered in PaLung cells. When describing these new results in the manuscript text, we will point out explicitly that the GSEA pathways chosen for display/discussion are not the absolute top pathways shown by GSEA, but among the top few metabolic gene sets and were selected by this protocol for display. In addition, we will also add supplemental material showing ES plots similar to the ones in Figure 1 for other similar gene sets that may be ranked higher in the GSEA ranking than the ones displayed in Figure 1.

3. In the Materials and methods section (lines 419-421), the authors mention, 'GSEA was run against the complete Gene ontology biological process (GO BP) gene set list (containing 18356 gene sets).' However, they narrow down the gene dataset for analysis (lines 136-138, 'we filtered our gene dataset to contain only genes listed under the Gene ontology category Cellular Metabolic Process (GO ID:0044237), resulting in a truncated list of 4794 genes.'). I'm concerned that this selective approach might introduce bias into the resultant pathways. Is this selective approach commonly employed in this type of analysis? And isn't there a need for adjustments to avoid potential bias?

We thank the reviewer for identifying this error on our side. It is true that in our initial analysis, we ‘filtered our gene dataset to contain only genes listed under the Gene ontology category Cellular Metabolic Process (GO ID:0044237), resulting in a truncated list of 4794 genes.’ Hence this line was included in the original version of the manuscript. However, owing to the same concerns raised by this reviewer, we later decided against this approach and ran the GSEA analysis using ALL genes detected in our RNAseq experiments (without filtering for metabolic genes). Figures 1C-D, and Supplementary Tables 2-3 were all generated from this GSEA analysis with the complete gene set list. However, it was an oversight on our part to not amend the manuscript text in lines 136-138. We propose to remove the corresponding lines in the revised manuscript.

4. The authors noted that the number of differentially expressed genes (DEGs) is quite high (6,247 out of 14,986) as per lines 134-135, stating that "The number of differentially expressed genes (6,247) was extremely high, suggesting that multiple pathways are differentially regulated between the two species." However, this large number of DEGs could indicate either an improper correction procedure or a need for a more stringent threshold. The authors should address this issue to avoid potential misinterpretation of the results.

We agree with this suggestion. The words “differentially expressed” are problematic in an inter-organism comparison. In our original submission, we never used the 6247 “differential expressed” genes as a pre-filter for downstream analysis, and our GSEA analyses used ALL genes common to both organisms. For the revised manuscript, we will refrain from using the term differentially expressed in the transcriptomic context. As suggested by reviewer 3, we will simply refer to them as genes that pass our cutoff threshold. In addition, we will also make our thresholds more stringent (p < 0.01 and |log fold change| > 1.5).

In terms of the correction procedure, in addition to the typical normalization performed, we also do correct for differences in transcript/gene length in different organisms as part of our analysis, and we believe this correction should suffice for inter-species comparison of RNAseq experiments.

5. In Figure 2B, the samples labeled as W1 and P1 appear to be outliers. This raises questions about the integrity of the sampling or analysis process. Please describe about this.

We would like to thank the reviewer for the comment and concern. Indeed, proteomics was done in pairs of PaLung and WI38 samples on different days, (P1, W1), (P2, W2), (P3, W3). It is possible that there might have been batch effects in the first pair that caused them to be outliers. Therefore, in addition to the main figure (n = 3), we will redo the proteomics analysis, including GSEA, after discarding the pair (P1, W1) and include this additional analysis in the supplementary figures.

6. Regarding the GSEA analysis of Figure 2, they are using the full set of GSEA. However, this reviewer is wondering if this is appropriate when analyzing mitochondrial fractions, as I believe using the entire GSEA set could introduce a bias. Is this a common approach? Shouldn't the authors be focusing on mitochondrial-related sets within the GSEA, and then determining the upregulated and downregulated pathways from there?

In response to this reviewer’s question, we propose to redo the GSEA analysis of the proteomics data using gene sets that are more specific for mitochondria. An example can be found in the Mitocarta database (https://www.broadinstitute.org/mitocarta/mitocarta30-inventory-mammalian-mitochondrial-proteins-and-pathways), which contains a gene set file (MitoPathways3.0.gmx). This gene set file is tailored for GSEA and contains 1136 mitochondrial genes sorted into 149 mitochondrial pathways/gene sets.

7. The authors describe in lines 195-197, "GSEA-flagged upregulation in OxPhos was driven mostly by the upregulation of Complex I subunits, for both the proteomic and transcriptomic data (Figure 2G, Supplementary Figure S1D)." However, within this analysis, the number of genes composing each subgroup of the mitochondrial Complexes are 44 for Complex I, 4 for Complex II, 10 for Complex III, and 19 for Complex IV (https://www.genenames.org/data/genegroup/#!/group/639). The authors mention that the genes of Complex I were dominant in the ETC, but, might this just be reflecting the original difference in the number of genes? As this reviewer believes this could have a significant impact on the authors' current claims, this reviewer suggest the authors to carefully reconsider this point, comparing the actual results with the proportion expected from the difference in gene numbers. (Even in Figure S1D, it appears to correlate with the number of genes: C1 39.3%, C3 10.7%, C4 10.7%, C2 3.5%)

Combined response for points 7 and 8 below.

8. As pointed out in Major Point 7, if the authors' claim of enrichment in Complex I is indeed due to the large number of genes included in the Complex I subgroup (https://www.genenames.org/data/genegroup/#!/group/639), can the assumption of High Complex I flux truly be considered valid? In that case, this constraints model would become inappropriate, and the validity of the inferred low or reverse activity of Complex II would be diminished. Therefore, a careful re-examination is desirable.

We interpret points 7 and 8 as having distinct but related concerns. (1) Is the enrichment of ETC just an artifact of Complex I having many subunits? (2) Is the ETC gene set enriched because of upregulation Complex I, or because of multiple mitochondrial complexes and it only appears to be dominated by the Complex I significance due to Complex I having many subunits? (3) Is the enrichment of Complex I just an artifact of Complex I having many subunits? Responses to these concerns are below.

In the first concern, the reviewers ask if it’s true that the proportion of observed Complex I genes (out of all possible Complex I genes) is larger than the proportion of observed Complex II, Complex III or Complex IV genes (compared to all possible Complex II, Complex III, and Complex IV genes), in the final GSEA core enrichment set. The reviewer suggested we compare the actual results versus the proportion expected from the difference in gene numbers. Therefore, we performed the following statistical analysis:

The null hypothesis is that the number of Complex I-IV genes observed in the core enrichment set match the frequencies of genes expected from the number of subunits in each complex. For transcriptomics, a chi-square test of goodness-of-fit for the observed counts of ETC genes in the core enrichment set vs expected counts based on gene proportions (Complex I: 44; Complex II: 4; Complex III: 10; Complex IV: 19) fails to reject the null hypothesis (that the observed counts are drawn from the expected frequencies) with a p-value of 0.6419 (chi-square statistic = 1.6779; dof = 3). The failure to reject the null hypothesis means that the actual results are similar to the expected frequencies by proportion.

This is no contradiction with our statements, because we do not claim that Complex I is preferentially enriched in PaLung in comparison to other ETC complexes. In fact, we do not make any claims or assertions about the other complexes of the ETC such as Complex II, Complex III and Complex IV from a transcriptomic or proteomic perspective.

To address the second concern, we will rephrase our text slightly, because we don’t mean to make a claim that Complex I is the only reason why the ETC gene set is enriched. Our claim is only that due to the absolute number of genes in Complex I (which is proportional to expected frequencies), significant enrichment of Complex I alone may be sufficient for GSEA to call the entire ETC gene set upregulated, without requiring a strong signal-to-noise ratio of enrichment due to the other complexes. We will modify the text in our manuscript to clarify this issue.

For the third concern, our claim of upregulated Complex I is substantiated by the fact that in proteomics, gene sets specific to Complex I are upregulated in PaLung. This observation is independent of other ETC Complex genes.

In summary, our claims are the following,

  • It is true that Complex I-specific gene sets are upregulated in PaLung cells. This underlies our assumption of higher flux through Complex I for flux modelling.

  • It is possible that these upregulated Complex I genes may be causing the broader ETC gene set to be flagged as upregulated.

  • It is not possible to infer from this analysis alone whether Complex II, Complex III and Complex IV are also up- or down-regulated, because individual gene sets for these complex activity/assembly were not flagged as up- or down-regulated in our analysis (unlike with Complex I).

9. (option, takes about 1-2 months). This reviewer believes that the authors' most important claim, concerning the high activity of Complex I and the low activity of Complex II, lacks strong evidence as no biochemical data of the activities of each mitochondrial complex are presented to substantiate this. Unless additional biochemical experimental data is provided, the assertions should be toned down. While the abstract mentions "complex II activity may be low or reversed," it is stated with certainty in line 108 of the introduction, "associated with the low or reverse activity of Complex II." Based on the present data, this reviewer believes that the claim remains speculative. Therefore, I suggest moderating the overall argument or adding the biochemical data. While the results from metabolomics are supportive, they do not serve as direct evidence.

We agree, and we thank the reviewer for noticing that certain locations have a higher strength of assertion. We will change the flagged line (Line 108) and moderate our assertion about possible low-to-reverse activity of Complex II and avoid over-interpreting the data.

10. Regarding Figure 5, the title of the figure states "lower antioxidant response", but it doesn't seem that the data in the figure actually shows a lower antioxidant response.

We thank the reviewer for noticing this discrepancy. We propose to modify the title of Figure 5 to “ROS and antioxidant system measurements in PaLung and WI-38 cells”.

11. In lines 109-110 of the Introduction, the authors state, "we confirmed our prediction of ischemic-like basal metabolism in PaLung cells by characterizing the response of bat cells to cellular stresses such as oxidative stress, nutrient deprivation, and a type of cell death related to ischemia, viz. ferroptosis." However, can the assertion that the cells are in an ischemic-like state be confirmed simply because they are resistant to several types of cellular stress?

We will moderate our language to clarify that resistance to nutrient deprivation and cellular stress are consistent with an ischemic-like state, but do not confirm an ischemic-like state. This change applies to both the results & Discussion sections.

Reviewer 1 Minor comments

1. The authors mention the use of cufflinks/Tophat for mapping/quantification. However, support for these software programs has ended and the creators of these programs themselves recommend using the successor programs. I recommend re-analysis using a more current pipeline (such as HISAT2/StringTie, STAR/RSEM, etc.). Furthermore, the transcriptomics section of the methods should also include the program used for cleaning and trimming.

We thank the reviewer for bringing our attention to this issue. For our revision plan, we propose to redo the transcriptomic analysis with the newer pipelines suggested by the reviewer and update our RNAseq results in the manuscript text, figures, and supplementary material as required.

2. As for the Oxygen Consumption Rate (OCR) data presented in Figure 2F, it makes sense that it's low at the basal level. However, it's perplexing that it is also low even under uncoupled conditions, especially considering the high energy demand associated with flight in this species. Could the authors provide their interpretation on this apparent contradiction?

We would like to thank the reviewer for the constructive and insightful comments. We also did not expect that bat cells would have a lower capacity than WI-38 cells under uncoupled conditions (which indicate maximal respiratory capacity). Given that flight is an energetically demanding activity (Maina JN 2000), bat cells are expected to have a high oxygen consumption capacity. Future research can test one possible interpretation, which is that this phenomenon might be cell-type dependent. For example, bat muscle cells, which possess numerous mitochondria and are the primary drivers of flight, might have a higher maximal respiratory capacity than human muscle cells. In contrast, bat fibroblasts, which might play a less important role in flight, might not require such a high capacity. Additionally, P. alecto can hibernate. To endure the challenging conditions of hibernation, bat fibroblasts may have adapted to reduce their oxygen intake and limit energy use. Because further analysis of different cell types will be essential to pursue these issues, we will incorporate them into the Discussion section. We hope this will be useful for future research to better understand the cell biology of bats.

Reference

Maina JN. What it takes to fly: the structural and functional respiratory refinements in birds and bats. J Exp Biol. 2000, 203:3045–3064, doi: 10.1242/jeb.203.20.3045

3. In line 156, the authors mention that 'Profiling detected a total of 1,469 proteins.' Please provide more details in the explanation. Specifically, does this total of 1,469 proteins represent a combined count from both humans and bats, or is this the number of proteins for which orthologs could be identified in both species, just like the authors did with the transcript results.

We will clarify that “The list of 1,469 proteins is composed of all proteins whose peptides were detected with high confidence in both species. There were no peptides detected in our experiment that were exclusively detected in high confidence in only one organism”. Further explanation in the methods section will indicate that orthologs for all 1469 proteins were identified using a combination of inParanoid and BLAST.

4. In Supplementary Table 4, only 127 mitochondrial proteins are listed out of the 405 proteins mentioned in "Of these 405 proteins, we identified 127 to be core mitochondrial proteins (lines 161-163)". As there is no explanation for this within Supplementary Table 4, it would be better to include one.

Thank you for this comment. In our revised manuscript, we shall update this table to contain data about all 405 mitochondrial proteins and not only the 127 proteins that exceeded our differential expression cutoffs.

5. In line 472, the phrase "GO BB gene set list" is used. Could this potentially be a typographical error, and should it instead be "GO BP gene set list"?

This was a typo. It should be the GO BP gene set list and will be corrected in the revised manuscript.

6. In the volcano plot of Figure S3B, it appears that the side with lower P/W values generally corresponds with lower p-values. I wonder if there might have been any oversight or mistake in the data analysis process that could explain this observation?

We thank the author for catching this puzzling asymmetry. It is true that in the volcano plot in S3B, many proteins that are high in WI-38 and low in bats (with a low P/W ratio) happen to fail statistical significance. To delve deeper into this issue, we extracted four subsets of the total set of 1469 proteins as follows based on their p-value (p) and log fold change (LFC).

  • Subset 1 (323 proteins): Low log(p) and low LFC (p > 0.05 and |LFC| < -1). This is the set of proteins that are highly enriched in the WI-38 sample but do not pass statistical significance.

  • Subset 2 (118 proteins): High log(p) and low LFC (p <= 0.05 and |LFC| < -1). This is the set of proteins that are highly enriched in the WI-38 sample and pass statistical significance.

  • Subset 3 (143 proteins): Low log(p) and high LFC (p > 0.05 and |LFC| > 1). This is the set of proteins that are highly enriched in the PaLung sample and do not pass statistical significance.

  • Subset 4 (289 proteins): High log(p) and high LFC (p <= 0.05 and |LFC| > 1). This is the set of proteins that are highly enriched in the PaLung sample and pass statistical significance.

For both PaLung and WI-38 samples and for each subset, we looked at the histogram of coefficient of variation (CV; standard deviation/mean) for the absolute abundance of each protein (See Author response image 1). Compared to the other histograms, we found that the CV histogram for the WI-38 samples in subset 1 had a more uniform distribution and included much higher values. Delving deeper into WI-38 subset 1, we identified 180 proteins that had a CV value greater than 0.7, suggesting that these proteins showed either high standard deviations or low mean abundance. Plotting the mean log(abundance) for the WI-38 samples across all four subsets, we found that subset 1 did not have lower abundances than the other subsets (Author response image 2). The abundance distribution for the 180 high-CV proteins also matched the abundance distribution for the rest of subset 1. All of this suggests that the observed asymmetry in the volcano plot is caused by a high standard deviation in 180 proteins in the WI-38 sample, which could be a batch and species-specific effect of metabolic regulation.

Author response image 1. Histograms of coefficient-of-variation (CV) for the four protein subsets (defined by p-value and log fold change).

Author response image 1.

Each row corresponds to a different subset indicated by the title above. Within each row, the blue histogram corresponds to the CV from the 3 PaLung samples, and the red histogram corresponds to the CV from the 3 WI-38 samples. The one major discrepancy is that the histogram for the WI-38 samples in subset 1 (Row one right panel), contains much higher values than the other 7 histograms.

Author response image 2. Histograms of log(abundance) for the four protein subsets (defined by p-value and log fold change) for WI-38 samples.

Author response image 2.

Each panel corresponds to a different subset indicated by the title above.

7. In lines 249-252, it is stated, "The low or negative flux values for Complex II in our PaLung simulations indicate that the electrons obtained from Complex I may accumulate at Complex II or potentially even get consumed by Complex II operating in reverse (bypassing the rest of the ETC) in PaLung cells." However, isn't the basic process of electron transfer done through Complex I-III-IV, independent of Complex II?

We thank the reviewer for raising this point. It is true that in conventional ETC, both Complex I and Complex II operate in parallel and feed electrons to Complex III and downstream. However, in cases of dysregulated ETC observed with the reverse activity of Complex II, it is possible that the electrons generated by Complex I do not proceed to Complex III and could be utilized by Complex II for its reverse activity (Bisbach et al., Cell reports 2020; Chouchani et al., Nature 2014). This phenomenon has been pointed out by earlier studies, two examples of which are shown below. In place of our earlier sentence, we will amend our manuscript text to say the following-

“During conventional ETC, both Complex I and Complex II operate in parallel and produce electrons that are shuttled downstream t to Complex III, Complex IV, and ATP synthase. However, prior work has documented an alternative in which the electrons obtained from Complex I can be consumed by Complex II operating in reverse, rather than traversing the rest of the ETC. This alternative was utilized by the computational models that showed low or reverse activity of Complex II.

8. Regarding Figure 4F, the authors state, 'PaLung cells displayed higher viability than WI-38 cells after glucose deprivation (Figure 4F).' However, in addition to the cell images, it would be beneficial to perform experimental quantification of cell death to provide more rigorous data. Additionally, the cells appear to be over-confluent, which might influence the results. Also, scale bars should be included in all photos, including Figure 6.

We are grateful for the valuable suggestions and feedback from the reviewer. For Figure 4F, we will repeat the experiments to quantify cell death at the same cell density as in Figure 6A and C. We will avoid the over-confluent conditions and include the scale bars in all the photos.

9. Regarding Figure 5B, it is stated that 'the expression levels of differentially expressed antioxidant genes' are shown, but it includes those that are not significant. It would be helpful if the authors could clarify how this gene set was selected.

Thank you for catching this omission in our explanation. The genes included in Figure 5B were only those that showed statistically significant differences. Asterisks are missing for three of the proteins. Since we will redo the transcriptomic analysis, the revised figure will have minor differences in the counts and TPM, and we will also correct the annotations in the updated figure. To address possible concerns about the set of antioxidant genes, we will plot all genes in the HALLMARK_REACTIVE_OXYGEN_SPECIES_PATHWAY set that are statistically significant in expression difference between PaLung and WI-38 cells.

10. Regarding Figure 6C, the values for total glutathione seem to significantly differ from those in Figure 5C. An explanation for this discrepancy would be appreciated to ensure the consistency and reliability of the data.

We thank the reviewers for the careful and critical reading of our manuscript. We apologize for our careless mistake. We noticed that the values in Figure 5C were normalized by protein amount, but Figure 6F was not. We will correct the labelling of the y-axis and replace Figure 6F with a graph showing the total glutathione normalized by protein amount (nmol/mg protein). After normalization, the values in Figure 6F are similar to those in Figure 5C.

Reviewer 2 Major comments

1. The authors compared a fibroblast cell line derived from adult bats with a human embryonic cell line. Please discuss whether mitochondrial metabolism in embryonic cells might be different and how it could have affected the obtained results. Please describe in more detail how the cells were established, what population doubling they were used at (both bat and human cells). Were the cells cultured in atmospheric oxygen or low-oxygen conditions. The exposure of cells to atmospheric oxygen might affect the many mitochondrial parameters measured in this study and could influence the main finding about ischemic-like state. Additionally, please mention in the limitations of the study that only biological n=1 was compared (since cells only from 1 individual per species was used in experimental groups), despite n=3 technical replicates.

We would like to thank the reviewer for the help in improving the quality of our work. We will make the necessary revisions to ensure that our Materials and methods section is clear and complete, and that we address the limitations of our study in the Discussion section.

Regarding the comparison of fibroblast cell lines from adult bats and human embryo stage, we acknowledge the potential differences in mitochondrial metabolism between these cell types. In the revised manuscript, we will discuss the differences in more detail, using the existing literature on mitochondrial metabolism.

We will describe the methods used to establish and propagate these cell lines, including the population doubling at which they were utilized in our experiments. We cultured them in a normoxic condition (not in a hypoxic condition). We will mention in the Discussion section that the exposure of cells to different oxygen levels could affect the metabolic phenotypes.

We appreciate the reviewer's comments that our study used cells from only one individual of each species, which is a major limitation. We will explicitly mention this in the Discussion section to ensure that readers are aware of the limitations of the sample size.

2. Reference genomes for bats are not as well annotated as for human. Downregulation of a pathway may result from some genes being excluded from the analysis because of poor annotation of the P. Alecto genome compared to human. The authors state: "Genes with counts per million (CPM) < 1 in more than 3 out of 6 samples were discarded from downstream analysis". So, if the gene was not annotated, was it assigned a zero value and discarded? Was it discarded if it was zero in one species (e.g. bat) or set to 0? If such genes were excluded, while in reality not being mapped, they could have skewed the pathway analysis.

The reviewer raises an important point here. It is true that reference genomes for less studied organisms like bats may not be as well annotated as other organisms. Indeed, the seeming down-regulation of a pathway in bats might be because the constituent genes were not annotated as well, not because they weren’t expressed as highly. For exactly this reason, we analysed pathways that were upregulated in bats, and made no conclusion about pathways downregulated in bats. We will make this more clear in our text.

For the question of discards, the reviewer’s interpretation is correct. Our transcriptomic comparison was performed only using genes/transcripts common to both organisms. If a gene was not detected in the bat experiment, it was removed from the comparison. We will clarify in the discussion that this discard would exclude genes that are unmapped in bat, and the pathway analysis might under-estimate bat-specific biology.

4. The major findings of this paper were based on the omic data, followed by some experimental validations. However, the quality of these omic data or the results are not solid enough to motivate the authors to validate these findings. For example, both of the GO terms enriched by the DEGs in Figure 1 are not the top terms as claimed by the authors (not even significant after multiple test correction). Also, even though the 2 GO terms in Figure 2 are quite significant, the expression pattern seems not very consistent among the replicates, which make the enrichments not so solid. This highlights an inconsistency among different omic datasets, which may generate some conflicting results. For example, the low level of metabolites from TCA cycle (Figure 4c) seems not consistent with the high level of TCA-related protein, as described in Figure 2c & d. For the purpose of improving the manuscript quality, the authors may have to evaluate the consistency among the multiple omic datasets or to optimize their bioinformatic pipeline to enhance the results.

Regarding the choice of GO terms, we agree this should be cleaner. When we redo our transcriptomic analysis with STAR, the results will be displayed using topmost gene sets according to the criteria in our response for Major Point 2 of Reviewer 1.

Regarding the consistency between different omic measurements, we believe it is possible to make inferences about pathways or gene sets when multiple -omic measurements agree at multiple levels of biology (metabolites / proteins / RNA). For example, when metabolites, proteins, and transcripts exhibit up-regulation of the same pathway, then we can infer with moderate confidence that the pathway has increased utilization or increased importance. However, the converse is not true. The absence of agreement between different levels of biology (metabolites / proteins / RNA) could be due to genuine biological complexities and/or errors of measurement/analysis. Many biological changes occur at one level of measurement without altering other levels of measurement, such as phosphorylation affecting protein degradation without affecting RNA. Any pathway can have up-regulation of one measure and down-regulation of another due to genuine regulatory mechanisms. Therefore, multi-omic agreement is a positive result, but multi-omics disagreement produces no result, and does not produce a problem or a contradictory result. Previous studies found a correlation < 0.3 between transcriptomic and proteomic levels in the same cells (Haider S and Pal R 2013, Gunawardana Y et al. 2013,Bathke J et al. 2019, Xu JY et al. 2020).Thus, it is no surprise that our data for bat and human ratios (P/W) have a correlation of 0.223 between the proteomic fold-change and the transcriptomic fold-change (see Author response image 3).

Author response image 3. Scatter plot of Log fold change of proteomic data vs Log fold change of transcriptomics data.

Author response image 3.

Each point represents a protein/gene that was detected in both proteomics and transcriptomics experiments. Log fold change was computed in each case as log2(PaLung /WI-38).

Reference:

Bathke, J., Konzer, A., Remes, B., McIntosh M., Klug G. Comparative analyses of the variation of the transcriptome and proteome of Rhodobacter sphaeroides throughout growth. BMC Genomics. 2019, 20:1: 358. https://doi.org/10.1186/s12864-019-5749-3.

Haider S, Pal R. Integrated analysis of transcriptomic and proteomic data. Curr Genomics. 2013, 14:2:91-110. doi: 10.2174/1389202911314020003.

Xu JY, Zhang C, Wang X, Zhai L, Ma Y, Mao Y, Qian K, Sun C, Liu Z, Jiang S, Wang M, Feng L, Zhao L, Liu P, Wang B, Zhao X, Xie H, Yang X, Zhao L, Chang Y, Jia J, Wang X, Zhang Y, Wang Y, Yang Y, Wu Z, Yang L, Liu B, Zhao T, Ren S, Sun A, Zhao Y, Ying W, Wang F, Wang G, Zhang Y, Cheng S, Qin J, Qian X, Wang Y, Li J, He F, Xiao T, Tan M. Integrative Proteomic Characterization of Human Lung Adenocarcinoma. Cell 2020, 182:1:245-261.e17. doi: 10.1016/j.cell.2020.05.043.

Gunawardana Y, Niranjan M, Bridging the gap between transcriptome and proteome measurements identifies post-translationally regulated genes, Bioinformatics 2013, 29:23: 3060–3066, https://doi.org/10.1093/bioinformatics/btt537

5. The dominant up-regulation of complex I in ETC is interesting and is the main finding of this paper. However, no experimental evidence was provided to prove the greater activity of Complex I, for example, metabolites changes. In addition, the genes encoding proteins belong to ETC complex I, II, III and IV vary a lot, with much more genes encoding complex I. Therefore, the author should consider the background gene number when they compare the up-regulated gene number differences in each complex. For example, a fisher-exact test could be done to see if complex I has significantly more genes been up-regulated than a random expectation.

We thank the reviewer for raising this issue. This is similar to reviewer 1’s concern about the proportion of ETC complexes raised in major points 7 and 8. As per our response earlier, we do not claim that Complex I genes are upregulated more than other complexes. A chi-square test shows that the proportion of ETC complex genes observed in the core enrichment set of the GSEA gene set is in keeping with the expected frequencies based on the number of genes in each complex. Our claim is only that due to the sheer number of genes in Complex I (even though it is in proportion with expected frequencies), significant enrichment of Complex I alone may be sufficient in most cases to call the entire ETC gene set upregulated, without requiring a strong signal-to-noise ratio of enrichment from the other complexes. We will modify our manuscript to clarify this, and we are grateful for feedback to improve our communication.

Reviewer 2 Minor comments

- The author may have to add the p value or FDR for each GSEA plot, even though some of the FDR are not significant. Also, it will be better to show the normalized enrichment score (NES) instead of the ES.

We thank the reviewer for this suggestion and will include the p value/ FDR for the gene sets displayed in the figures.

- The gene set name in several supplementary tables contains many '%' characters and those needs to be removed.

We will amend this in the supplementary material of our revised submission.

- In Line 302, "…combined with the earlier findings of downregulated OxPhos expression and low OCR, we conclude…". If my understanding is right, the authors only mentioned the up-regulation of Oxphos expression, instead of down-regulation. This sentence may need to be clarified.

We will amend this text in our revised submission.

Reviewer 3 Major comments

1. The authors state:

"We then set the lower bound of the PaLung Complex I reaction flux to a value equal to 70% of its theoretical maximum. Similarly, we set the upper bound of the WI-38 Complex I reaction at a value equal to 30% of its theoretical maximum value. This ensured that the PaLung model would have higher flux through the Complex I reaction, in comparison to the WI-38 model."

How do the results hold with different thresholds ? Are these findings robust with e.g. in ranges between 10 to 50% (90-50%) (instead of only 30% and 70%). Furthermore, the histogram figures doesnt seem to reflect a 70% of maximum lower bound for complex I (threshold at a value of 30 seems like extremity of tail).

The reviewer raises a good question, whether our simulation results depend on the thresholds we used. We propose to perform additional flux sampling with thresholds proposed by the reviewer in the range of 10-50%. To help readers interpret the modelling already shown, we will point out that the output flux histograms are the result of combined constraints on Complex I and oxygen intake, so the feasible ranges can be narrower than the flux constraint on Complex I. For the simulations suggested by the reviewer, we warn that using the suggested 10-50% thresholds might generate a conflict with oxygen intake rates and could result in infeasible scenarios. (In that case, the set of possible flux configurations would be the empty set, and would not be plotted in the revised manuscript.)

2. Number of differentially expressed genes is extremely high because such cutoffs are not really meaningful given the comparison between two organisms. No need to refer to the 6247 above cutoff as differentially regulated genes (see: https://elevanth.org/blog/2023/07/17/none-of-the-above/ and https://daniel-saunders-phil.github.io/imagination_machine/posts/if-none-of-the-above-then-what/ for pointers toward current best practice in biological statistics). Enough to simply note that 6247 are above the cutoffs, which suggest a drastic (and expected) difference in expression profiles between the two organisms.

We thank the reviewer for this suggestion. We will refrain from saying “differentially expressed” when comparing different species, and will instead refer to the differences as genes that pass the cutoff threshold.

3. Please highlight the RNA and proteomic analysis assumption and present results within those boundaries (e.g. how are the transcript matched between human and bat, the use of human gene ontologies, etc…). Are the human GO set definitions relevant in bat (it is a common practice with mice and rats, are bats close ?)?

We thank the reviewer for raising this concern. Unlike well studied organisms like human and mice, few bioinformatics resources are readily available for P. alecto. Hence in certain steps in our analyses, we used human-derived resources for inter-species comparison. In our revised manuscript, we will reiterate these methodology choices in the discussion :

  • Use of Human GO terms and gene set definitions for inter-species GSEA

  • Using only transcripts found in both organisms for transcriptomic analyses

  • Mapping of proteomics peptides to human and P. alecto proteins.

  • Use of human derived metabolic mitochondrial models for flux sampling (Mitocore)

4. Are oxphos and hypoxia responses the most extreme pathway scores in the GSEA ? Instead of barcode plots that are generally not a very useful use of figure space, use Figure 1C to show the top e.g.20 (positive and negative) pathway scores so that we can see how much those two actually stand out. Same for the proteomic analysis. Also, need to show an unbiased side by side comparison of the pathway enrichments for RNA and proteomic, the reported results in main text and figures are too cherry picked to be of interest as they stand.

We prefer to provide the full table in supplementary materials because truncating after the top 20 entries would yield too many transcriptomic differences that are irrelevant to metabolism or mitochondria, and therefore impossible to corroborate with the mitochondrial proteomic data. The current figures will be enhanced by adding p-values, but we prefer not to delete the GSEA plots because the p-values don’t provide enough information about the strength of evidence.

5. Finally, and very importantly, please upload ALL the code used for the analysis, with instructions to run it and all the required inputs and source files. The computational analysis is only as credible as it is easy to reproduce.

We apologize if the reviewer found difficulties in running our code (which was uploaded at https://github.com/nsuhasj/PalungWI38FluxSim). We did identify an oversight in the upload, which is the need for better documentation about installing externally available code libraries. We will also perform a blind download on a fresh workstation to confirm that our github repository provides all necessary supporting files.

Reviewer 3 Minor comments

Introduce GeTMM, what are its key specificities ?

We propose to add more details about GeTMM in our revised manuscript.

Figure 1C code bar plot useless, simply report ES and NES and pathway absolute rank in text.

We acknowledge the reviewer’s comment, but as mentioned in our earlier response, we would prefer to retain the bar plot over the list of gene sets with ES/NES.

Report Foldchange/p-value/rank of complex-I members and other genes of interest for the narrative of the paper.

We will add supplemental material indicating these values.

Reviewer 2 Major comments

3. All conclusions are based on high-throughput data, however it is accepted that some validation should be provided. Please provide qPCR or WB (if good antibodies are available) validation for several most significantly differentially expressed genes supporting the pathways identified in Figure 2 (preferably supporting the conclusions about Complexes I/II).

We appreciate the reviewer's valuable suggestions. As the reviewer is aware, it is challenging to find human antibodies to detect bat proteins. Furthermore, for a fair comparison, human antibodies should detect bat proteins with the same affinity as they do human proteins. We aligned the Complex I proteins from humans and bats and found that their homology was relatively low, which suggests that it would be difficult to validate our findings using Western blotting.

Unfortunately, the same challenges apply to validation by quantitative PCR (qPCR). Comparing gene expression by qPCR between different species can be difficult due to several factors, including genetic differences, primer specificity, and the choice of reference genes.

Different species can have variations in their gene sequences, which can affect the binding of qPCR primers and lead to differences in amplification efficiency. Therefore, primers must be designed to target conserved regions of the gene of interest to ensure specificity. However, finding such regions that are conserved between species can be difficult when the gene sequences are not well conserved. In addition, even if the primer annealing region is conserved, amplification efficiency may not be the same if the sequences of the entire amplicon have some variations. As mentioned above, the sequences of the genes in Complex I between humans and bats are not very well conserved.

Nevertheless, we selected eight genes in Complex I that were upregulated in PaLung bat cell lines in our RNAseq or proteomics analysis of mitochondrial proteins, and tested the primers. These genes were NDUFA3, NDUFA7, NDUFA10, NDUFA13, NDUFB2, NDUFB9, NDUFS2, and NDUFV2. The primer sequences were 100% matched to human sequences, but had one or two mismatches to bat sequences. Although we were able to amplify the human fragments well, we were unable to amplify the bat fragments of these genes efficiently or at all. Furthermore, we could not amplify the GAPDH fragments for both human and bat cell lines at comparable levels, even though the primer sequences for GAPDH were 100% matched to both human and bat sequences. Possible explanations for this include variation in GAPDH expression levels between the cell lines, differences in GAPDH amplification efficiency during PCR, or a combination of both factors, even though GAPDH is one of the most commonly used reference genes for qPCR analysis.

In this situation, we believe that RNAseq provides a fair comparison between two different species, as it represents the percentage of expression of a given gene relative to the total mRNA expression. Therefore, rather than attempting to evaluate our findings by Western blot or qPCR, we decided to moderate our conclusions and avoid over-interpreting the data, as suggested by Reviewer 1 in Major Comment 9.

6. If the main findings of this paper can be further confirmed by additional experiments or data, it will be a very nice paper. This could be a potential mechanism that bats used to switch metabolism modes between two metabolic extremes: flight and hibernation, which require high and low energy. However, the usage of only the lung fibroblasts of human and bat may limit the ability of generalizing this 'ischemic-like state' of ETC in most of the bats tissue/organs. While I agree what the authors mentioned in the Discussion section, that to extend to primary cells of other species can help generalize this finding, studying the metabolism state of different cell type of bats (e.g., muscle cells responsible for flight; myocytes and neurons for hibernation) probably can provide more insights into the evolution of various interesting phenotypes of bats.

We would like to thank the reviewer for the insightful comments. We agree that extending the current study to different bat cell types, such as myocytes and neurons, could provide a more comprehensive understanding of metabolic adaptations in bat. This valuable suggestion is beyond the capacity of our laboratory resources and the timeline of this work (i.e., acquisition of primary tissues and generation of cell lines). Our current study provides the first comprehensive comparison of bat and human, to generate interest for subsequent work. We sincerely apologise for this limitation and appreciate the reviewer's understanding. We will include an additional discussion indicating that our work, characterizing differences in TCA cycle, OxPhos and ROS responses in PaLung, acts as a stepping stone for further research that can deepen our understanding of bat metabolism and physiology.

Reviewer 2 Minor comments

- How did mitochondrial DNA content per cell compared between the two species? Could the results be affected by the number and size of the mitochondria per cell in each species? An indirect measurement of mitochondrial DNA yield in the fractionation experiment would be the total DNA amount that was obtained in mitochondrial fractions per cell lysed.

We appreciate the reviewer’s feedback and thoughtful suggestions. The questions about mitochondrial DNA content per cell, and the potential impact of mitochondrial number and size, are indeed relevant to studies of mitochondrial function and metabolism. However, we believe these measurements may not be essential in the context of our study. Regarding the comparison of mitochondrial DNA content between the two species, the genes we observed to be differentially expressed in the mitochondria of bats and humans were encoded by nuclear DNA. As such, measuring mitochondrial DNA content would not significantly alter our main findings.

We acknowledge that our results could potentially be affected by mitochondrial morphology, including the number, size, and fragmentation status of mitochondria. One common approach to evaluating mitochondrial morphology is the use of fluorescence dyes specific to mitochondria, in addition to electron microscopy. However, we found that PaLung bat cells can export most of the staining dyes used to visualize mitochondria, which makes it difficult to accurately assess these parameters. The strong export activity in PaLung cells is due to the high expression of ABCB1 transporter (Koh et al., Nat Commun 2019).

In summary, dyes for assessing mitochondrial morphology have had different behaviors and affinities across species, so we do not wish to assert any conclusions that may be generated through these techniques. We request your understanding that such technical difficulties have prevented us from performing a morphological evaluation of mitochondria to further elucidate our findings. We will include this in the Discussion section.

Reference:

Koh J, Itahana Y, Mendenhall IH, Low D, Soh EXY, Guo AK, Chionh YT, Wang LF, Itahana K. ABCB1 protects bat cells from DNA damage induced by genotoxic compounds. Nat Commun. 2019 Jun 27;10(1):2820. doi: 10.1038/s41467-019-10495-4.

Reviewer Comments

1) The code and source data availability are still lack luster. None of the required files to run the only script that they provide are available as of now, neither are instructions on how to run it.

We agree that our first Github upload was insufficient. We have now updated the manuscript with the link to a new Github repository. This repository contains additional files (both data files and scripts) and a README file with instructions on how to run the scripts (and install necessary software). We have also verified that it is possible to run the scripts by performing the installation on a fresh workstation. Our contact information is provided at the repository, and we welcome feedback, including anonymous email.

2) They didn't address my concern about figure 2B constraint looking inconsistent between what is said in method and what is shown on the plot.

We would first like to clarify if the Reviewer might be referring to Figure 3C and not 2B? Figure 2B is a proteomic heatmap and does not involve any constraints.

We interpret the reviewer’s question as an extension of the following query from the original reviewer comments at ReviewCommons

“The authors state:

"We then set the lower bound of the PaLung Complex I reaction flux to a value equal to 70% of its theoretical maximum. Similarly, we set the upper bound of the WI-38 Complex I reaction at a value equal to 30% of its theoretical maximum value. This ensured that the PaLung model would have higher flux through the Complex I reaction, in comparison to the WI-38 model."

How do the results hold with different thresholds ? Are these findings robust with e.g. in ranges between 10 to 50% (90-50%) (instead of only 30% and 70%). Furthermore, the histogram figures doesn’t seem to reflect a 70% of maximum lower bound for complex I (threshold at a value of 30 seems like extremity of tail).”

In our ReviewCommons response, we said, “To help readers interpret the modelling already shown, we will point out that the output flux histograms are the result of combined constraints on Complex I and oxygen intake, so the feasible ranges can be narrower than the flux constraint on Complex I.” That statement may have been insufficient, so we explain in more detail below.

The histograms in figure 3C show the feasible flux ranges for Complex I and Oxygen intake reactions, after constraints have been placed on both reactions. Because of the highly interconnected nature of metabolic networks, placing constraints on one reaction might indirectly limit the feasible flux space of another reaction, which is what we observe here in the case of Complex I (CI) and Mitochondrial Oxygen transport (O2) reactions. To illustrate this, we performed flux sampling simulations under four conditions:

  • Unconstrained control where both CI and O2 reactions can take any feasible flux value.

  • Where only CI fluxes are constrained to be 30-70, i.e., the CI flux of the bat model has its lower bound set at 70% of its maximum possible value, while that of the human model has its upper bound set at 30% of its maximum possible value.

  • Where only O2 fluxes are constrained to be 30-70, i.e., the O2 flux of the bat model has an upper bound set at 30% of its maximum possible value, while that of the human model has its lower bound set at 70% of its maximum possible value.

  • Fully constrained model where both reactions have been constrained as in Figure 3C of the main manuscript.

It can be seen from Author response image 4 that constraining only CI or only the O2 reaction flux also automatically constrains the other reaction (more p`ronounced in the bat model).

Author response image 4. Flux histograms showing the feasible flux distributions for the Complex I (CI) and the mitochondrial oxygen transport (O2) reactions in the P (PaLung) and W (WI-38) metabolic models.

Author response image 4.

Each column corresponds to setting constraints as per one of the four experiments described above.

In our updated manuscript, we also include Appendix 3 to illustrate the setting of constraints in more detail.

Reviewer:

3) As of now, the narrative they present for the choice of focus on the two metabolite pathways they consider from the RNA and proteomic GSEA analysis does not hold. The two pathways they focus on are blended in the middle of hundreds of other pathways that have as much or more evidence of deregulation, but they do not provide any rationale as to why they specifically chose to focus on these two.

Thank you for the opportunity to explain. Below is a detailed explanation of how we chose to focus on the pathways displayed in our main text figures 1 and 2.

Transcriptomics GSEA (Figure 1)

It is true that the two pathways we display “Respiratory electron transport…” and “Cellular response to hypoxia” are not the top differentially regulated pathways and are blended among other altered pathways. Because we seek metabolic differences, we chose gene sets for display as follows:

  • We first looked only at pathways upregulated in the PaLung samples. This is because pathways downregulated in PaLung could be artifacts due to incomplete/partial annotation of the PaLung genome compared to the human genome. Our updated Discussion section explains this rationale.

  • We then selected gene sets that satisfy the FDR < 0.25 threshold commonly used for GSEA analysis. Justification for why a higher FDR threshold is used for GSEA can be found on the GSEA FAQ website at the following URL https://software.broadinstitute.org/cancer/software/gsea/wiki/index.php/FAQ

  • Next, we selected gene sets that had a nominal enrichment score with absolute value >= 1

  • Finally, we manually searched this truncated list of gene sets for pathways related to metabolism, especially “primary metabolism” (central carbon metabolism).

The unfiltered GSEA list is provided as a supplement, and Author response table 1 is a list of the 136 gene sets that passed the first 3 steps of the filtering process, prior to any manual search. We have highlighted in orange the 21 gene sets that we identified as most relevant to metabolism. Note that most of these 21 metabolism-relevant gene sets are NOT part of central carbon metabolism.

Author response table 1.

NAME SIZE NES FDR
L13A-MEDIATED TRANSLATIONAL SILENCING OF CERULOPLASMIN EXPRESSION REACTOME DATABASE ID RELEASE 71 156827 102 -1.86459 0
CAP-DEPENDENT TRANSLATION INITIATION REACTOME DATABASE ID RELEASE 71 72737 110 -1.85211 0
TRANSLATIONAL INITIATION GOBP GO:0006413 114 -1.8446 0
REGULATION OF EXPRESSION OF SLITS AND ROBOS REACTOME DATABASE ID RELEASE 71 9010553 153 -1.84266 0
NONSENSE MEDIATED DECAY (NMD) ENHANCED BY THE EXON JUNCTION COMPLEX (EJC) REACTOME R-HSA-975957.1 107 -1.83796 0
GTP HYDROLYSIS AND JOINING OF THE 60S RIBOSOMAL SUBUNIT REACTOME R-HSA-72706.2 103 -1.82911 0
NUCLEAR-TRANSCRIBED MRNA CATABOLIC PROCESS GOBP GO:0000956 179 -1.82134 0
MRNA CATABOLIC PROCESS GOBP GO:0006402 191 -1.81824 0
EUKARYOTIC TRANSLATION INITIATION REACTOME DATABASE ID RELEASE 71 72613 110 -1.81618 0
EUKARYOTIC TRANSLATION ELONGATION REACTOME R-HSA-156842.2 85 -1.81611 0
SRP-DEPENDENT COTRANSLATIONAL PROTEIN TARGETING TO MEMBRANE REACTOME R-HSA-1799339.2 103 -1.81578 0
SELENOCYSTEINE SYNTHESIS REACTOME DATABASE ID RELEASE 71 2408557 84 -1.81445 0
NONSENSE-MEDIATED DECAY (NMD) REACTOME R-HSA-927802.2 107 -1.81251 0
TRANSLATION REACTOME DATABASE ID RELEASE 71 72766 278 -1.8121 0
RNA CATABOLIC PROCESS GOBP GO:0006401 214 -1.81042 0
PEPTIDE BIOSYNTHETIC PROCESS GOBP GO:0043043 313 -1.8098 0
FORMATION OF A POOL OF FREE 40S SUBUNITS REACTOME DATABASE ID RELEASE 71 72689 92 -1.8098 0
INFLUENZA LIFE CYCLE REACTOME DATABASE ID RELEASE 71 168255 129 -1.80935 0
RESPONSE OF EIF2AK4 (GCN2) TO AMINO ACID DEFICIENCY REACTOME DATABASE ID RELEASE 71 9633012 92 -1.80856 0
VIRAL MRNA TRANSLATION REACTOME DATABASE ID RELEASE 71 192823 81 -1.80841 0
PEPTIDE CHAIN ELONGATION REACTOME R-HSA-156902.2 81 -1.80743 0
NONSENSE MEDIATED DECAY (NMD) INDEPENDENT OF THE EXON JUNCTION COMPLEX (EJC) REACTOME R-HSA-975956.1 87 -1.80499 0
INFLUENZA VIRAL RNA TRANSCRIPTION AND REPLICATION REACTOME DATABASE ID RELEASE 71 168273 121 -1.80381 0
PROTEIN LOCALIZATION TO ENDOPLASMIC RETICULUM GOBP GO:0070972 119 -1.80193 0
NUCLEAR-TRANSCRIBED MRNA CATABOLIC PROCESS, NONSENSE-MEDIATED DECAY GOBP GO:0000184 107 -1.79958 0
SIGNALING BY ROBO RECEPTORS REACTOME R-HSA-376176.4 193 -1.79657 0
INFLUENZA INFECTION REACTOME R-HSA-168254.2 139 -1.79422 0
TRANSLATION GOBP GO:0006412 296 -1.79232 0
EUKARYOTIC TRANSLATION TERMINATION REACTOME R-HSA-72764.4 85 -1.79057 0
SELENOAMINO ACID METABOLISM REACTOME DATABASE ID RELEASE 71 2408522 105 -1.7747 0
MAJOR PATHWAY OF RRNA PROCESSING IN THE NUCLEOLUS AND CYTOSOL REACTOME R-HSA-6791226.3 170 -1.77321 0
ESTABLISHMENT OF PROTEIN LOCALIZATION TO ENDOPLASMIC RETICULUM GOBP GO:0072599 100 -1.76931 0
COTRANSLATIONAL PROTEIN TARGETING TO MEMBRANE GOBP GO:0006613 90 -1.7682 0
RRNA PROCESSING IN THE NUCLEUS AND CYTOSOL REACTOME R-HSA-8868773.3 180 -1.76363 2.64E-05
PROTEIN TARGETING TO MEMBRANE GOBP GO:0006612 134 -1.76038 2.57E-05
CYTOPLASMIC RIBOSOMAL PROTEINS WIKIPATHWAYS_20191210 WP477 HOMO SAPIENS 82 -1.75783 4.97E-05
PROTEIN TARGETING TO ER GOBP GO:0045047 97 -1.75446 7.25E-05
VIRAL GENE EXPRESSION GOBP GO:0019080 122 -1.75351 7.06E-05
SRP-DEPENDENT COTRANSLATIONAL PROTEIN TARGETING TO MEMBRANE GOBP GO:0006614 85 -1.75031 6.88E-05
VIRAL TRANSCRIPTION GOBP GO:0019083 105 -1.74701 6.71E-05
ACTIVATION OF THE MRNA UPON BINDING OF THE CAP-BINDING COMPLEX AND EIFS, AND SUBSEQUENT BINDING TO 43S REACTOME R-HSA-72662.3 56 -1.74344 8.72E-05
AMIDE BIOSYNTHETIC PROCESS GOBP GO:0043604 383 -1.74084 8.51E-05
NUCLEOBASE-CONTAINING COMPOUND CATABOLIC PROCESS GOBP GO:0034655 322 -1.73596 1.04E-04
CYTOPLASMIC TRANSLATION GOBP GO:0002181 54 -1.73436 1.02E-04
RRNA PROCESSING REACTOME DATABASE ID RELEASE 71 72312 189 -1.73364 9.96E-05
TRANSLATION INITIATION COMPLEX FORMATION REACTOME DATABASE ID RELEASE 71 72649 55 -1.73183 9.75E-05
RIBOSOMAL SCANNING AND START CODON RECOGNITION REACTOME R-HSA-72702.3 55 -1.71426 2.11E-04
PROTEIN TARGETING GOBP GO:0006605 283 -1.70904 2.62E-04
FORMATION OF THE TERNARY COMPLEX, AND SUBSEQUENTLY, THE 43S COMPLEX REACTOME DATABASE ID RELEASE 71 72695 48 -1.70325 4.03E-04
CELLULAR NITROGEN COMPOUND CATABOLIC PROCESS GOBP GO:0044270 345 -1.69784 4.66E-04
ESTABLISHMENT OF PROTEIN LOCALIZATION TO ORGANELLE GOBP GO:0072594 325 -1.69391 6.33E-04
AROMATIC COMPOUND CATABOLIC PROCESS GOBP GO:0019439 347 -1.6925 6.90E-04
PEPTIDE METABOLIC PROCESS GOBP GO:0006518 392 -1.69196 6.94E-04
CELLULAR RESPONSES TO STRESS REACTOME DATABASE ID RELEASE 71 2262752 456 -1.67591 0.001561
HETEROCYCLE CATABOLIC PROCESS GOBP GO:0046700 343 -1.66849 0.00217
ESTABLISHMENT OF PROTEIN LOCALIZATION TO MEMBRANE GOBP GO:0090150 215 -1.66714 0.00226
CELLULAR RESPONSES TO EXTERNAL STIMULI REACTOME DATABASE ID RELEASE 71 8953897 459 -1.66708 0.002236
ORGANIC CYCLIC COMPOUND CATABOLIC PROCESS GOBP GO:1901361 365 -1.64869 0.004763
CALNEXIN CALRETICULIN CYCLE REACTOME R-HSA-901042.2 23 -1.60474 0.022972
N-GLYCAN TRIMMING IN THE ER AND CALNEXIN CALRETICULIN CYCLE REACTOME DATABASE ID RELEASE 71 532668 32 -1.59981 0.026341
RIBOSOMAL LARGE SUBUNIT BIOGENESIS GOBP GO:0042273 64 -1.58547 0.041181
OXYGEN-DEPENDENT PROLINE HYDROXYLATION OF HYPOXIA-INDUCIBLE FACTOR Α REACTOME DATABASE ID RELEASE 71 1234176 61 -1.58214 0.045084
ER QUALITY CONTROL COMPARTMENT (ERQC) REACTOME DATABASE ID RELEASE 71 901032 18 -1.57559 0.054216
RIBOSOME ASSEMBLY GOBP GO:0042255 49 -1.56563 0.071338
AMINO ACID AND DERIVATIVE METABOLISM REACTOME R-HSA-71291.6 282 -1.56487 0.071759
CITRIC ACID CYCLE (TCA CYCLE) REACTOME DATABASE ID RELEASE 71 71403 22 -1.56128 0.078243
TRANSLATION FACTORS WIKIPATHWAYS_20191210 WP107 HOMO SAPIENS 48 -1.55985 0.079794
REGULATION OF TP53 DEGRADATION REACTOME R-HSA-6804757.1 31 -1.55649 0.086393
VIRAL PROCESS GOBP GO:0016032 259 -1.5553 0.08777
SIGNALING BY FGFR4 REACTOME R-HSA-5654743.2 27 -1.55424 0.088823
REGULATION OF CALCIUM-MEDIATED SIGNALING GOBP GO:0050848 47 -1.55084 0.095609
NEGATIVE REGULATION OF G0 TO G1 TRANSITION GOBP GO:0070317 36 -1.54615 0.106841
CYCLIN D ASSOCIATED EVENTS IN G1 REACTOME R-HSA-69231.7 40 -1.54534 0.107785
G1 PHASE REACTOME DATABASE ID RELEASE 71 69236 40 -1.54481 0.107951
INSULIN PROCESSING REACTOME R-HSA-264876.2 20 -1.54226 0.114012
ERROR-PRONE TRANSLESION SYNTHESIS GOBP GO:0042276 19 -1.53741 0.126991
TRANSLESION SYNTHESIS BY POLH REACTOME R-HSA-110320.1 18 -1.53506 0.132554
SMOOTH MUSCLE CONTRACTION REACTOME R-HSA-445355.3 29 -1.53357 0.136112
CELLULAR RESPONSE TO HYPOXIA REACTOME R-HSA-1234174.2 69 -1.53084 0.143644
INTERSPECIES INTERACTION BETWEEN ORGANISMS GOBP GO:0044419 322 -1.53011 0.144136
DUAL INCISION IN TC-NER REACTOME R-HSA-6782135.1 62 -1.53011 0.142368
TCA CYCLE (AKA KREBS OR CITRIC ACID CYCLE) WIKIPATHWAYS_20191210 WP78 HOMO SAPIENS 18 -1.52732 0.149964
SYMBIONT PROCESS GOBP GO:0044403 316 -1.52634 0.151093
REGULATION OF TP53 EXPRESSION AND DEGRADATION REACTOME R-HSA-6806003.1 32 -1.5262 0.149924
ERROR-FREE TRANSLESION SYNTHESIS GOBP GO:0070987 19 -1.52572 0.149766
TRANSCRIPTIONAL REGULATION BY E2F6 REACTOME DATABASE ID RELEASE 71 8953750 34 -1.52515 0.150193
MITOPHAGY REACTOME DATABASE ID RELEASE 71 5205647 25 -1.52402 0.152274
GAP-FILLING DNA REPAIR SYNTHESIS AND LIGATION IN TC-NER REACTOME DATABASE ID RELEASE 71 6782210 62 -1.52125 0.161169
ER-PHAGOSOME PATHWAY REACTOME DATABASE ID RELEASE 71 1236974 69 -1.52107 0.159943
SYNTHESIS OF ACTIVE UBIQUITIN: ROLES OF E1 AND E2 ENZYMES REACTOME R-HSA-8866652.2 29 -1.52027 0.160681
FCERI MEDIATED NF-ΚB ACTIVATION REACTOME R-HSA-2871837.2 71 -1.51881 0.164231
SIGNALING BY Fgfr1 REACTOME DATABASE ID RELEASE 71 5654736 32 -1.51836 0.164082
CENTRAL NERVOUS SYSTEM NEURON DEVELOPMENT GOBP GO:0021954 29 -1.51808 0.163292
REGULATION OF TP53 ACTIVITY THROUGH METHYLATION REACTOME DATABASE ID RELEASE 71 6804760 17 -1.51684 0.166234
AUF1 (HNRNP D0) BINDS AND DESTABILIZES MRNA REACTOME DATABASE ID RELEASE 71 450408 50 -1.51593 0.167798
B CELL ACTIVATION REACTOME DATABASE ID RELEASE 71 983705 92 -1.5138 0.173984
CYTOPLASMIC PATTERN RECOGNITION RECEPTOR SIGNALING PATHWAY GOBP GO:0002753 32 -1.51365 0.172791
IKK COMPLEX RECRUITMENT MEDIATED BY RIP1 REACTOME DATABASE ID RELEASE 71 937041 18 -1.51332 0.172421
VIRION ASSEMBLY GOBP GO:0019068 35 -1.51249 0.173743
ENDOSOMAL SORTING COMPLEX REQUIRED FOR TRANSPORT (ESCRT) REACTOME R-HSA-917729.1 27 -1.5117 0.174884
INFECTIOUS DISEASE REACTOME DATABASE ID RELEASE 71 5663205 379 -1.51083 0.176457
CHONDROITIN SULFATE METABOLIC PROCESS GOBP GO:0030204 26 -1.50909 0.181727
PROTEIN LOCALIZATION TO MEMBRANE GOBP GO:0072657 357 -1.5088 0.18113
I-KAPPAB KINASE/NF-KAPPAB SIGNALING GOBP GO:0007249 51 -1.50817 0.181788
TICAM1, RIP1-MEDIATED IKK COMPLEX RECRUITMENT REACTOME R-HSA-168927.3 17 -1.50722 0.183755
RIBOSOME BIOGENESIS GOBP GO:0042254 227 -1.50672 0.183798
VESICLE-MEDIATED TRANSPORT BETWEEN ENDOSOMAL COMPARTMENTS GOBP GO:0098927 21 -1.50447 0.190799
DEGRADATION OF GLI2 BY THE PROTEASOME REACTOME R-HSA-5610783.1 56 -1.50387 0.191333
CITRATE METABOLIC PROCESS GOBP GO:0006101 29 -1.50382 0.189784
ATP METABOLIC PROCESS GOBP GO:0046034 136 -1.50016 0.202786
PRADER-WILLI AND ANGELMAN SYNDROME WIKIPATHWAYS_20191210 WP3998 HOMO SAPIENS 31 -1.49851 0.20778
MYD88-INDEPENDENT TOLL-LIKE RECEPTOR SIGNALING PATHWAY GOBP GO:0002756 27 -1.49841 0.206341
ANTIGEN PROCESSING AND PRESENTATION OF PEPTIDE ANTIGEN VIA MHC CLASS I GOBP GO:0002474 80 -1.49762 0.207835
STABILIZATION OF P53 REACTOME R-HSA-69541.5 53 -1.49515 0.216427
DOWNSTREAM TCR SIGNALING REACTOME R-HSA-202424.3 76 -1.49442 0.217854
THE ROLE OF GTSE1 IN G2 M PROGRESSION AFTER G2 CHECKPOINT REACTOME DATABASE ID RELEASE 71 8852276 57 -1.49062 0.233181
ABC TRANSPORTER DISORDERS REACTOME DATABASE ID RELEASE 71 5619084 64 -1.49037 0.232277
GAP-FILLING DNA REPAIR SYNTHESIS AND LIGATION IN GG-NER REACTOME R-HSA-5696397.1 24 -1.48948 0.234394
TNFR2 NON-CANONICAL NF-ΚB PATHWAY REACTOME R-HSA-5668541.2 80 -1.48934 0.233057
CLEC7A (DECTIN-1) SIGNALING REACTOME R-HSA-5607764.1 89 -1.4891 0.232317
REGULATION OF G0 TO G1 TRANSITION GOBP GO:0070316 38 -1.48792 0.235624
NEGATIVE REGULATION OF FGFR4 SIGNALING REACTOME DATABASE ID RELEASE 71 5654733 19 -1.48738 0.236268
VIF-MEDIATED DEGRADATION OF APOBEC3G REACTOME DATABASE ID RELEASE 71 180585 50 -1.48711 0.235552
G1 S DNA DAMAGE CHECKPOINTS REACTOME R-HSA-69615.2 63 -1.48708 0.233782
GLI3 IS PROCESSED TO GLI3R BY THE PROTEASOME REACTOME R-HSA-5610785.1 56 -1.48693 0.232652
RESPIRATORY ELECTRON TRANSPORT, ATP SYNTHESIS BY CHEMIOSMOTIC COUPLING, AND HEAT PRODUCTION BY UNCOUPLING PROTEINS. REACTOME R-HSA-163200.1 107 -1.48672 0.231731
REGULATION OF TP53 ACTIVITY THROUGH PHOSPHORYLATION REACTOME DATABASE ID RELEASE 71 6804756 85 -1.48642 0.231313
BUDDING AND MATURATION OF HIV VIRION REACTOME DATABASE ID RELEASE 71 162588 24 -1.48594 0.231623
SIGNALING BY FGFR REACTOME R-HSA-190236.2 61 -1.48574 0.230753
PHOSPHODIESTERASES IN NEURONAL FUNCTION WIKIPATHWAYS_20191210 WP4222 HOMO SAPIENS 29 -1.48396 0.237066
P53-DEPENDENT G1 DNA DAMAGE RESPONSE REACTOME DATABASE ID RELEASE 71 69563 61 -1.48333 0.237846
PINK PARKIN MEDIATED MITOPHAGY REACTOME R-HSA-5205685.2 20 -1.48228 0.24089
MFAP5 EFFECT ON PERMEABILITY AND MOTILITY OF ENDOTHELIAL CELLS VIA CYTOSKELETON REARRANGEMENT WIKIPATHWAYS_20191210 WP4560 HOMO SAPIENS 16 -1.48205 0.240132
FC EPSILON RECEPTOR (FCERI) SIGNALING REACTOME DATABASE ID RELEASE 71 2454202 108 -1.48203 0.238453
OXIDATIVE STRESS INDUCED SENESCENCE REACTOME R-HSA-2559580.4 65 -1.48099 0.241159

Only 21 gene sets pass the filtering for relevance and significance (from a starting count of 18000+ gene sets). Of these 21 gene sets, 5 pathways related to TCA cycle/electron transport chain (ETC) and 3 gene sets related to hypoxia and stress response. Many of the other pathways pertain to anabolism/catabolism or secondary metabolism (e.g., RNA catabolism, translation, amide metabolic process, seleno amino acid metabolism, chondroitin sulfate metabolism etc). This is why we chose to highlight the TCA/ETC gene set and the hypoxia gene set in Figure 1.

Although the selected gene sets are not the topmost altered metabolic pathways in our list, they are still significantly altered, according to FDR and NES. Furthermore, by belonging to central carbon metabolism, they provide greater insight into fundamental metabolic differences between the two species compared.

Our manuscript now includes the following text in the transcriptomics result section.

“Since the P. alecto genome is less fully annotated than the human genome, pathways with incomplete annotation may be incorrectly predicted to be downregulated in PaLung cells. Hence, we only studied differentially regulated pathways that were upregulated in PaLung. When we filtered PaLung-upregulated gene sets for significance (indicated by FDR<0.25 and normalized enrichment score |NES| > 1), and for relevance to metabolism, only 21 gene sets remained. Many were relevant to secondary metabolism or anabolic/catabolic housekeeping, five were related to the TCA cycle and electron transport (including “ATP synthesis by chemiosmotic coupling, respiratory electron transport, and heat production by uncoupling proteins”) and three were related to hypoxic stress (such as “Cellular response to hypoxia” in the Reactome Pathway Database).”

Proteomics GSEA (Figure 2)

Following similar logic as with the transcriptomics GSEA (upregulated in PaLung, FDR < 0.25, |NES| > = 1), resulted in a total of 118 gene sets. The top 20 gene sets in this list (Author response table 2) contained 8 gene sets that pertained to metabolism (highlighted in orange).

Author response table 2.

NAME SIZE NES FDR
MUSCLE CONTRACTION GOBP GO:0006936 35 -2.29 0.008034
NICOTINIC ACETYLCHOLINE RECEPTOR SIGNALING PATHWAY PANTHER PATHWAY P00044 15 -2.17947 0.018238
REGULATION OF CELL JUNCTION ASSEMBLY GOBP GO:1901888 15 -2.17782 0.012159
THE CITRIC ACID (TCA) CYCLE AND RESPIRATORY ELECTRON TRANSPORT REACTOME R-HSA-1428517.1 64 -2.13697 0.015138
HALLMARK_OXIDATIVE_PHOSPHORYLATION MSIGDB_C2 HALLMARK_OXIDATIVE_PHOSPHORYLATION 91 -2.12375 0.013852
COLLAGEN FORMATION REACTOME DATABASE ID RELEASE 71 1474290 31 -2.11295 0.013374
MUSCLE SYSTEM PROCESS GOBP GO:0003012 40 -2.07522 0.017172
EPH-EPHRIN SIGNALING REACTOME DATABASE ID RELEASE 71 2682334 30 -2.06553 0.016753
ACTIN FILAMENT-BASED MOVEMENT GOBP GO:0030048 17 -2.05773 0.017006
RESPIRATORY ELECTRON TRANSPORT REACTOME R-HSA-611105.3 32 -2.0515 0.016612
RHO GTPASES ACTIVATE PKNS REACTOME DATABASE ID RELEASE 71 5625740 17 -2.0504 0.015369
COLLAGEN BIOSYNTHESIS AND MODIFYING ENZYMES REACTOME R-HSA-1650814.3 25 -2.03023 0.018264
RESPIRATORY ELECTRON TRANSPORT, ATP SYNTHESIS BY CHEMIOSMOTIC COUPLING, AND HEAT PRODUCTION BY UNCOUPLING PROTEINS. REACTOME R-HSA-163200.1 32 -2.0253 0.017967
COMPLEX I BIOGENESIS REACTOME R-HSA-6799198.1 16 -2.02379 0.016892
MITOCHONDRIAL RESPIRATORY CHAIN COMPLEX I ASSEMBLY GOBP GO:0032981 16 -2.01705 0.01673
INTEGRIN SIGNALLING PATHWAY PANTHER PATHWAY P00034 44 -2.00758 0.017239
NADH DEHYDROGENASE COMPLEX ASSEMBLY GOBP GO:0010257 16 -1.99919 0.017902
MITOCHONDRIAL ATP SYNTHESIS COUPLED ELECTRON TRANSPORT GOBP GO:0042775 26 -1.99497 0.01756
ACTOMYOSIN STRUCTURE ORGANIZATION GOBP GO:0031032 21 -1.98512 0.018578
KERATINIZATION GOBP GO:0031424 15 -1.96272 0.024099

All 8 pathways belonged to the TCA cycle/ETC, within which 3 correspond to Complex I of ETC, which is why we display these gene sets in Figure 2 of the main text.

An additional response to the Reviewer’s question is also now provided in Appendix 2, which addresses Reviewer 1’s request (Major comment 6). This supplement re-computes our proteomics GSEA using a mitochondria-only geneset list (Mitocarta 3.0; URL: https://www.broadinstitute.org/mitocarta/mitocarta30-inventory-mammalian-mitochondrial-proteins-and-pathways), instead of the full GO BP gene set list. Results from this new GSEA show that OxPhos is the top upregulated pathway in the bat samples. Complex I is also among the top 5 gene sets found upregulated in bat proteomic samples. Author response table 3 shows all the gene sets upregulated in the bat samples, with FDR value < 0.25.

Author response table 3. Gene sets enriched in phenotype P (3 samples).

GS follow link to MSigDB GS DETAILS SIZE ES NES NOM p-val FOR q-val FWER p-val
OXPHOS SUBUNITS Details. 34 -0.55 -2.19 o.ooo 0.002 0.001
2 OXPHOS Details. 41 -0.53 -2.17 o.ooo 0.001 0.001
3 CARBOHYDRATE METABOLISM Details. 36 -0.50 -2.00 o.ooo 0.005 0.009
4 TRANSLATION Details. 16 -0.62 -1.99 0.002 0.005 0.011
5 COMPLEX I Details. 17 -0.60 -1.97 0.004 0.005 0.013
6 Cl SUBUNITS Details. 15 -0.62 -1.89 0.004 0.007 0.024
7 FATTY ACID OXIDATION Details. 20 -0.55 -1.86 0.008 0.007 0.030
8 METALS AND COFACTORS Details. 30 -0.47 1.79 0.006 0.011 0.051
9 METABOLISM Details. 153 -0.34 1.78 0.002 0.011 0.058
10 AMINO ACID METABOLISM Details. 33 -0.46 1.76 0.011 0.011 0.066
11 MITOCHONDRIAL CENTRAL DOGMA Details. -0.45 -1.61 0.012 0.030 o. 181
12 TCA CYCLE Details. 15 -0.48 -1.51 0.052 0.053 0.310
13 LIPID METABOLISM Details. 43 -0.36 -1.48 0.032 0.057 0.359
14 PROTEIN IMPORT SORTING AND HOMEOSTASIS Details. 24 -0.34 -1.20 0.205 0.223 0.846

Associated Data

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

    Data Citations

    1. Koh J, Irving A, Itahana Y, Lee Y, Itahana K, Wang L, Suhas Jagannathan N, Tucker-Kellogg L. 2024. RNAseq comparison of lung fibroblasts from Pteropus alecto (PaLung cell line) and Homo sapiens (WI-38 cell line) NCBI Gene Expression Omnibus. GSE215934
    2. Itahana K. 2024. Multi-omic analysis of bat versus human fibroblasts reveals altered central metabolism. jPOST repository. JPST001821 [DOI] [PMC free article] [PubMed]
    3. Koh J, Jagannathan MS, Sobota RM, Tucker-Kellogg L, Itahana Y, Itahana K. 2024. Comparison of human and bat metabolism. ProteomeXchange. PXD043121
    4. Koh J, Jagannathan NS, Tucker-Kellogg L, Itahana Y, Itahana K. 2024. Metabolomics comparison of lung fibroblasts from Pteropus alecto and Homo sapiens. Metabolomics Workbench. ST002743

    Supplementary Materials

    Supplementary file 1. Genes that pass our differential expression cutoffs (false discovery rate [FDR] < 0.05; |log fold change| > 1) in PaLung vs WI-38 samples from whole-cell transcriptomics data.

    Differential expression analysis was performed using the DESeq2 pipeline. Fold changes are indicated as PaLung/WI-38.

    elife-94007-supp1.xlsx (657.9KB, xlsx)
    Supplementary file 2. Biological pathways upregulated in PaLung cells from gene set enrichment analysis (GSEA) of transcriptomics data.

    GSEA was performed on the transcriptomics data using (PI) value as a metric. The table below lists the pathways detected as upregulated in PaLung cells (compared to WI-38 cells) and associated enrichment metrics.

    elife-94007-supp2.xlsx (245.8KB, xlsx)
    Supplementary file 3. Biological pathways upregulated in WI-38 cells from gene set enrichment analysis (GSEA) of transcriptomics data.

    GSEA was performed on the transcriptomics data using (PI) value as a metric. The table below lists the pathways detected as upregulated in WI-38 cells (compared to PaLung cells) and associated enrichment metrics.

    elife-94007-supp3.xlsx (158.8KB, xlsx)
    Supplementary file 4. Differentially expressed (DE) mitochondrial proteins in PaLung vs WI-38 samples from mitochondrial proteomics data.

    405 DE proteins were first identified using a Student’s t-test on median-corrected protein abundances from the mitochondrial samples of PaLung and WI-38. Of the 405 DE proteins, 127 were identified to be core mitochondrial proteins (as defined by MitoCarta and IMPI datasets) and are listed in this sheet. Fold changes are indicated as PaLung/WI-38.

    elife-94007-supp4.xlsx (14.2KB, xlsx)
    Supplementary file 5. Biological pathways upregulated in PaLung cells from gene set enrichment analysis (GSEA) of proteomics data.

    GSEA was performed on the proteomics data using protein abundances as input. The table below lists the pathways detected as upregulated in PaLung cells (compared to WI-38 cells) and associated enrichment metrics.

    elife-94007-supp5.xlsx (56.1KB, xlsx)
    Supplementary file 6. Biological pathways upregulated in WI-38 cells from gene set enrichment analysis (GSEA) of proteomics data.

    GSEA was performed on the proteomics data using protein abundances as input. The table below lists the pathways detected as upregulated in WI-38 cells (compared to PaLung cells) and associated enrichment metrics.

    elife-94007-supp6.xlsx (81KB, xlsx)
    Supplementary file 7. Metabolic model for PaLung cells.

    A metabolic flux model was constructed for the central carbon metabolism of PaLung cells by overlaying proteomic and transcriptomic information onto the existing mitocore model from literature.

    elife-94007-supp7.xlsx (57.3KB, xlsx)
    Supplementary file 8. Metabolic model for WI-38 cells.

    A metabolic flux model was constructed for the central carbon metabolism of WI-38 cells by overlaying proteomic and transcriptomic information onto the existing mitocore model from literature.

    elife-94007-supp8.xlsx (60.4KB, xlsx)
    Supplementary file 9. Flux sampling results comparing flux distributions in the constrained PaLung and WI-38 models.

    Flux sampling was performed with 5000 flux vectors for the PaLung and WI-38 metabolic models each. The flux histograms for each reaction were compared across the two models and the following statistics were extracted from the histograms.

    elife-94007-supp9.xlsx (67.8KB, xlsx)
    Supplementary file 10. Absolute metabolite quantification in PaLung and WI-38 cells.

    Absolute concentrations of metabolites detected in PaLung and WI-38 cells by Human Metabolome Technologies (HMT).

    elife-94007-supp10.xlsx (40.7KB, xlsx)
    MDAR checklist

    Data Availability Statement

    Transcriptomic data are deposited in the NCBI GEO database (GSE215934). Proteomic data are deposited in the ProteomeXchange database (PXD043121) and in the jPOST repository (JPST001821). Metabolomics data are uploaded to the Metabolomics Workbench database (ST002743). The Matlab script used for flux sampling can be found at here (copy archived at Jagannathan, 2023). Other data supporting the findings of this study are available within the article and its supplementary materials.

    The following datasets were generated:

    Koh J, Irving A, Itahana Y, Lee Y, Itahana K, Wang L, Suhas Jagannathan N, Tucker-Kellogg L. 2024. RNAseq comparison of lung fibroblasts from Pteropus alecto (PaLung cell line) and Homo sapiens (WI-38 cell line) NCBI Gene Expression Omnibus. GSE215934

    Itahana K. 2024. Multi-omic analysis of bat versus human fibroblasts reveals altered central metabolism. jPOST repository. JPST001821

    Koh J, Jagannathan MS, Sobota RM, Tucker-Kellogg L, Itahana Y, Itahana K. 2024. Comparison of human and bat metabolism. ProteomeXchange. PXD043121

    Koh J, Jagannathan NS, Tucker-Kellogg L, Itahana Y, Itahana K. 2024. Metabolomics comparison of lung fibroblasts from Pteropus alecto and Homo sapiens. Metabolomics Workbench. ST002743


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

    RESOURCES