Abstract
Insight on the underlying mechanisms of aging will advance our ability to extend healthspan, treat age-related pathology, and improve quality of life. Multiple genetic and pharmacological manipulations extend longevity in different species, yet monotherapy may be relatively inefficient and we have limited data on the effect of combined intervention. Here we summarize interactions between age-related pathways and discuss strategies to simultaneously retard these in different organisms. In some cases, combined manipulations additively increase their impact on common hallmarks of aging and lifespan, suggesting they quantitatively participate within the same pathway. In other cases, interactions affect different hallmarks, suggesting their joint manipulation may independently maximize their effects on lifespan and healthy aging. While most interaction studies have been conducted with invertebrates and show varying levels of translatability, the conservation of pro-longevity pathways offer an opportunity to identify “druggable” targets relevant to multiple human age-associated pathologies.
Keywords: aging, longevity, healthspan, combined interventions, systems biology, complexity
BRIEF SUMMARY
This review provides an overview of combinatorial approaches used to extend lifespan across species. The authors identify frameworks that can be used for analyzing interactions between mechanisms of aging with the goal of finding translatable interventions.
I. A framework for interaction analysis in aging research
There are two goals in aging research: deciphering the mechanisms of aging, and discovering interventions that delay its manifestations. Although many single genetic and pharmacological manipulations are known to slow aging 1,2, it is unlikely that monotherapy will be effective because aging is an outcome of pleiotropic processes. Instead, combined interventions will interact and likely be more effective than any alone. Interaction analysis measures the phenotypic effects of two (or more) simultaneous aging interventions. The effects are measured as manifestations of aging, including the various hallmarks, inflammaging, specific age-associated diseases, age-acquired frailty, and age-dependent mortality. The manipulations include genetic alterations, pharmacology, environmental conditions such as diet and temperature, and physiological/surgical procedures. A substantial body of interaction analyses has developed over recent years, in which each study focuses on one biological specific. Here, we begin to assemble these data in a common framework.
Types of interventions
Perhaps the oldest intervention in aging research involves diet, originally applied to rodents through caloric and protein restriction. More recent studies also include macro-nutrient ratio, specific amino acid content, control of mealtime, or administration of enzymes that degrade specific nutrients 3. These interventions are now applied to many animal models and to humans. Diet interventions often are combined with genetic manipulations.
Among these combinations, genetic mutations of the insulin/IGF pathway provide a foundation for manipulative studies of aging in C. elegans and Drosophila. Current technologies have expanded possibilities for genetic interventions through targeted single nucleotide substitution of endogenous genes, tissue and temporal control of gene products, targeted gene ablation, RNA depletion, and more. The genetic approach is central to discovering the basic mechanisms of aging where a key feature of the approach is to study the interaction between two genetic manipulations or between a gene and an extrinsic intervention such as with diet or drugs.
Pharmacological interventions 4 are emerging as a powerful analytical approach, which is greatly enhanced when the work establishes a mechanistic basis. Well-studied examples, each with caveats, include rapamycin, spermidine, vitamin E, metformin, NAD+ metabolites and curcumin 4. Their hypothesized mechanisms of action span the range of molecular discovery in aging research. In each case, we can understand the mechanism by which such drugs slow aging by using genetic manipulations to determine what molecules and pathways are required for them to extend lifespan.
Topography of interactions
Interaction analysis can be improved when we consider how manifestations of aging are organized. These topographies identify the relationship between experimental interventions and their phenotypic effects, and ways to interpret mechanism.
One approach builds a hierarchy for aging from the perspective of high-throughput-omics 5 (Figure 1a). In this framework, tiers reflect the flow of biological information from genome to transcriptome, to proteome, to metabolome and ultimately to the aging phenome - the set of all cellular, tissue, organ and organismal age-associated phenotypes. Within each tier, nodes describe signaling and functional networks. In this approach, interaction analysis manipulates multiple features either within the same level or among two levels. We measure the phenome as the outcome. ‘Same level’ interventions may involve the transcriptome tier where we overexpress pro-longevity transcription factors FOXO and NRF2, or the metabolome tier where we feed animals various NAD metabolites. Interventions across tiers may include when long-lived eat-2 mutants (genome level) are supplemented with an H2S donor FW1256 (metabolome level). One challenge for this topography is the way one level can impact many nodes at the next level. For example, when transcription factors impact the expression of many proteins, it becomes difficult to assign causal direction within and between -omics levels. Hierarchical structures of aging need to also include feedback, which is not readily represented in the simple -omics topography. For example, the universal methyl donor, S-adenosylmethionine (metabolome tier) provides methyl groups to molecules belonging to all other tiers, while each of these layers subsequently feedback upon the methionine cycle and impact SAM 6. Similarly, glucose can cause non-enzymatic modification detected at the genome and the proteome, while events at these tiers feedback to the metabolome.
Figure 1. Hierarchical frameworks for aging interaction analysis.
a. Systems biology considers high-throughput-omics as a set of tiers that contain functional and signaling networks within levels. The effects of manipulations of features within an -omic tier or between two tiers can be observed at the level of the age-phenome, the set of all cellular, tissue, organ and organismal age-associated phenotypes. This approach can map aging-related interactions among levels, but is complicated when interventions impact multiple nodes upon the next level. Figure modified from 5. b. The ‘nominal aging’ hierarchy of Golubev 8. Asymmetry in chemical dynamics provides the underlying basis of aging. We observe increasingly complex, manifestations that are built upon processes of lower interactions. Hallmarks of Aging (bolded and underlined) are observed as specific molecular, cellular, and physiological manifestations of aging. By considering manipulations of processes at the cellular level via genetic or pharmacological studies, interactions among Hallmarks generate emergent manifestations of aging such as morbidity, frailty and mortality.
A second approach builds hierarchy based on “manifestations of aging”, which includes the so-called Hallmarks of Aging 7,8 (Figure 1b). In this view, aging originates from the nonenzymatic interactions between transient metabolites such as glucose, reactive oxygen species, and carbonyls, with slow-turnover, information-dense macromolecules of metabolic control such as DNA, proteins and ribozymes. These chemical processes yield molecular manifestations such as protein damage and misfolding, DNA damage and mutation, and mitochondrial damage. Interactions between these molecular manifestations propagate to produce Hallmarks of Aging (stem cell exhaustion, epigenetic alterations, loss of proteostasis, etc.), which subsequently reverberate to physiological manifestations (sarcopenia, cancer, inflammation) and then to emergent properties such as reserve depletion, functional decline, and accelerated mortality rate. Multiple interventions can be made within and between these levels. We measure how their interaction affects manifestations at similar or higher levels. For example, mitochondrial fusion (mitochondrial dysfunction hallmark) may be altered with age by simultaneously manipulating protein misfolding and nuclear DNA damage (molecular manifestations) 9. Manipulating protein damage and clearing senescent cells (features of molecular and hallmark manifestations) may alter sarcopenia (physiological manifestation) 10,11.
A key feature of this aging-manifestation topography is how different levels interact and involve feedback 8. Responses at higher levels intrinsically affect those at lower levels such that relationships across levels cannot be inferred only from analyzing interactions at one tier. For example, altered intercellular communication (Hallmark of Aging) can produce compensatory effects upon both mitochondrial and DNA damage responses, which themselves interact to affect various Hallmarks. Such feedback provides avenues for homeostasis and may also explain why transient stresses such as a burst of ROS produces hormesis. The interconnected structure produces a systems biology that is unique for each species and may determine the propensity for perceived Hallmarks framework, physiological manifestations, or nominal features such as the mortality rate doubling time characteristic of each species.
Relative to these frameworks, interaction analysis asks: can the simultaneous use of two interventions improve healthy aging more than either intervention alone? What does the impact of combined interventions reveal about the underlying mechanisms of their action? The first question is often addressed empirically: combine two interventions and see if the outcome exceeds the benefit of each individual case. The second question provides a way to predict productive combinations and builds mechanistic insight that will drive future discovery. Table 1 surveys a range of published intervention interactions including manipulations of nutrient sensing, oxygen sensing, mitochondria, redox regulation, lipid metabolism, reproduction, and microbiota.
Table 1. Dual interventions for extension of lifespan.
Studies that have manipulated two nodes within an individual hallmark of aging or two different hallmarks and shown an additive, partial, or synergistic effect on lifespan. Select studies of interest are summarized here, with extent of effect added where relevant. Note that lifespan effects were included in the table only when explicitly specified in the original publication as mean or median lifespan extension.
Organism | Intervention I | Signaling pathway (Hallmark of aging) I | % lifespan increase (Intervention I) | Intervention II | Signaling pathway (Hallmark of aging) II | % lifespan increase (Intervention II) | % lifespan increase (combination) | Reference |
---|---|---|---|---|---|---|---|---|
C. elegans | daf-2(e1370) | Deregulated nutrient sensing | 169% | rsks-1(ok1255) | Deregulated nutrient sensing | 20% | 454% | 20 |
C. elegans | daf-2(e1370) | Deregulated nutrient sensing | 111% | Rapamy cin (TOR inhibitor) | Deregulated nutrient sensing | 26% | 187% | 20 |
C. elegans | daf-2(e1370) | Deregulated nutrient sensing | LET-363/CeT OR | Deregulated nutrient sensing | 20 | |||
C. elegans | cyc-2.1 | Mitochondrial dysfunction | 64% | rsks-1 | Deregulated nutrient sensing | 14% | 54% | 21 |
C. elegans | cyc-2.1 | Mitochondrial dysfunction | 64% | daf-2 | Deregulated nutrient sensing | 114% | 252% | 21 |
C. elegans | RPS-15 | Deregulated nutrient sensing | eat-2 | Deregulated nutrient sensing | 22 | |||
C. elegans | RPS-22 | Deregulated nutrient sensing | eat-2 | Deregulated nutrient sensing | 22 | |||
C. elegans | rsks-1/S6K | Deregulated nutrient sensing | eat-2 | Deregulated nutrient sensing | 22 | |||
C. elegans | eIF2β | Deregulated nutrient sensing | eat-2 | Deregulated nutrient sensing | 22 | |||
C. elegans | eIF4G | Deregulated nutrient sensing | eat-2 | Deregulated nutrient sensing | 22 | |||
C. elegans | TOR/let-363 | Deregulated nutrient sensing | S6K/rsks-1(sv31) | Deregulated nutrient sensing | 22 | |||
C. elegans | TOR/let-363 | Deregulated nutrient sensing | eIF4E/ife - 2(ok306) | Deregulated nutrient sensing | 22 | |||
C. elegans | isp-1(qm150) | Mitochondrial dysfunction | rps-15 | Deregulated nutrient sensing | 22 | |||
C. elegans | isp-1(qm150) | Mitochondrial dysfunction | rps-22 | Deregulated nutrient sensing | 22 | |||
C. elegans | isp-1(qm150) | Mitochondrial dysfunction | S6K/rsks-1 | Deregulated nutrient sensing | 22 | |||
C. elegans | DR (limiting bacterial food) | Deregulated nutrient sensing | 37% | daf-2(e1370) | Deregulated nutrient sensing | 105% | 205% | 118 |
D. melanogaster | Rapamycin | Deregulated nutrient sensing | 10% | Chico heterozy gous flies | Deregulated nutrient sensing | 7% | 20% | 36 |
D. melanogaster | Rapamycin | Deregulated nutrient sensing | DR | Deregulated nutrient sensing | 36 | |||
D. melanogaster | Rapamycin | Deregulated nutrient sensing | 12–14% | Wortmannin | Deregulated nutrient sensing | 8% | 2.4-23.4% | 98 |
C. elegans | isp-1 | Mitochondrial dysfunction | nuo-6 | Mitochondrial dysfunction | 44 | |||
C. elegans | mrps-5 | Mitochondrial dysfunction | 18% | eat-3 | Mitochondrial dysfunction | 12% | 55% | 47 |
C. elegans | mrps-5 | Mitochondrial dysfunction | 18% | fzo-1 | Mitochondrial dysfunction | 7% | 36% | 47 |
C. elegans | drp-1 | Mitochondrial dysfunction | fzo-1 | Mitochondrial dysfunction | 47 | |||
C. elegans | T21C9.1/MICS-1 | Mitochondrial dysfunction | 54% | ATAD-3 | Mitochondrial dysfunction | 29% | 83% | 119 |
C. elegans | mtcu-2 | Mitochondrial dysfunction | mttu-1 | Mitochondrial dysfunction | 50% | 48 | ||
C. elegans | gas-1 | Mitochondrial dysfunction | T26A5.3 | Mitochondrial dysfunction | 45 | |||
C. elegans | gas-1 | Mitochondrial dysfunction | clk-1(e2519) | Mitochondrial dysfunction | 45 | |||
Rotifera | Trolox | Antioxidant | beta-carotene | Antioxidant | 10% | 60 | ||
Rotifera | Trolox | Antioxidant | L-carnosine | Antioxidant | 13% | 60 | ||
Rotifera | Trolox | Antioxidant | N-acetyl cysteine | Antioxidant | 16% | 60 | ||
Rotifera | Trolox | Antioxidant | EUK-8 | Antioxidant | 14% | 60 | ||
Rotifera | N-acetyl cysteine | Antioxidant | quercetin | Antioxidant | 14.5% | 60 | ||
Rotifera | EUK-8 | Antioxidant | L-carnosine | Antioxidant | 10% | 60 | ||
Rotifera | EUK-8 | Antioxidant | indole-3-propionic acid | Antioxidant | 14% | 60 | ||
C. elegans | ule-4 | Uterinelumen-expressed proteins | ule-5 | Uterine lumen-expressed proteins | 15% | 120 | ||
C. elegans | daf2 | Deregulated nutrient sensing | 100% | Germline removal | Reproduction | ~60% | ~300% | 50,51 |
C. elegans | daf-2(e1370) | Deregulated nutrient sensing | 253% | daf12(m20) | Nuclear receptor for dafachronic acid | −27% | 387% | 57 |
C. elegans | daf-2(e1370) | Deregulated nutrient sensing | daf-9 | Reproduction | 121 | |||
C. elegans | daf-2 | Deregulated nutrient sensing | 92% | sul-2 | Steroid hormone production | 32% | 124% | 58 |
C. elegans | daf-10(m79) | Cilia formation | 34% | sul-2 | Steroid hormone production | 32% | 89% | 58 |
C. elegans | daf-2(e1370) | Deregulated nutrient sensing | 106% | Pyrroloq uinoline quinone (PQQ) | antioxidant | 29% | 148% | 62 |
C. elegans | age-1(hx546) | Deregulated nutrient sensing | 36% | Pyrroloq uinoline quinone (PQQ) | antioxidant | 36% | 79% | 62 |
C. elegans | eat-2 | Deregulated nutrient sensing | 22% | Pyrroloq uinoline quinone (PQQ) | antioxidant | 26% | 42% | 62 |
C. elegans | hif-1(ia4) | Oxygen sensing | 18% | Pyrroloq uinoline quinone (PQQ) | antioxidant | 20% | 30% | 62 |
D. melanogaster | chico 1 | Deregulated nutrient sensing | Rhodiola rosea (RS) | antioxidant | 9–14% | 67 | ||
D. melanogaster | DR | Deregulated nutrient sensing | Rhodiola rosea (RS) | antioxidant | 13–36% | 67 | ||
C. elegans | sod-2 | Mitochondrial function | clk-1 | Mitochondrial function | 165% | 68 | ||
C. elegans | sod-2 | Mitochondrial function | eat-2 | Deregulated nutrient sensing | 68 | |||
C. elegans | sod-2 | Mitochondrial function | glp-1DR | Deregulated nutrient sensing | 68 | |||
Rotifera | Rapamycin | TOR inhibitor | 16% | SP600125 | JNK inhibitor | 20% | 33% | 122,123 |
Mice | dasatinib | inhibitor of multiple tyrosine kinases | quercetin | a natural flavonol | 36% | 79 | ||
Mice | metformin | No effect | rapamycin | mTORC1 inhibitor | 23% | 83 | ||
Mice | Rapamycin | mTORC1 inhibitor | acarbose | Alpha-glucosidase inhibitor | 28–34% | 86 | ||
Mice | Simvastatin | cholesterol biosynthesis inhibitor | ramipril | Angiotensin system inhibitor | 9% | 91 | ||
Mice | Ames dwarf mice | DR | Deregulated nutrient sensing | 28 | ||||
C. elegans | ethosuximide | anticonvulsant | eat-2(ad465) | Deregulated nutrient sensing | 68,92 | |||
C. elegans | ethosuximide | anticonvulsant | 13% | osm-3 | Sensory neuron function | 21% | 32% | 92 |
C. elegans | ethosuximide | Anticonvulsant | tax-4 | Sensory neuron function | 92 | |||
C. elegans | ethosuximide | Anticonvulsant | 13% | unc-31 | neurotransmission | 37% | 64% | 92 |
C. elegans | ethosuximide | Anticonvulsant | 13% | unc-64 | neurotransmission | 35% | 46% | 92 |
C. elegans | ethosuximide | Anticonvulsant | 17% | aex-3 | neurotransmission | 1 6% | 31% | 92 |
C. elegans | ethosuximide | anticonvulsant | 17% | daf-2 | Deregulated nutrient sensing | 107% | 135% | 92 |
Interactions among age-interventions are studied through epistasis analysis and gene-by-environment analysis. Epistasis analysis is a genetic approach in which the combined effect of two loci are compared to their independent effects 12. It is deceptively challenging to design and interpret these approaches. As Gems et al. highlighted, designs are compromised when 1) the interventions do not fully alter the targeted mechanism; 2) the interventions affect aging through multiple mechanisms; 3) interventions have different age-dependent effects on the age-trait; 4) suppressor mutants are deleterious, and 5) reaction norms are based on too few environmental levels. To address these issues, Gems et al. recommend we use null alleles rather than hypomorphs (often not an option), use more than one allele or transgene, and use alleles with few pleiotropic effects. For environmental interventions, Gems et al. suggested testing multiple treatment levels to identify the dose that produces the maximum lifespan. All these designs require appropriate statistical analysis and interpretation 13.
A further question asks whether the interaction observed in a model organism is universal for all organisms. Many examples exist where a manipulation is beneficial in one animal yet causes a devastating disease in humans. For instance, ribosomal downregulation may extend lifespan in worms but leads to ribosomopathy in humans 14; loss of OPA-1 (eat-3 in worms) leads to lifespan extension in worms but optic atrophy in humans 15; SOD-1 loss, which increases lifespan in worms, leads to ALS in humans 16. This problem is compounded in interaction analysis where one component negates the positive effects of others.
In the following we sample the space of published interaction analyses to illustrate these themes and to provide a foundation for further analyses of the complexity.
II. Interactions within Hallmarks
The Hallmarks of Aging are proposed categories of universal cellular and physiological aging 7,17. This expanding system identifies underlying molecular manifestations that impact complex phenotypes including morbidity and mortality. We illustrate Hallmark interactions by focusing on nutrient sensing and mitochondrial function, with additional cases compiled in Supplementary Table 1.
Interactions in Nutrient Sensing Hallmarks
Insulin/insulin-like growth factor 1 (IGF-1) signaling (IIS) and the target of rapamycin (TOR) (Figure 3) are conserved nutrient-sensing pathways shown to regulate lifespan 18,19. In C. elegans, double mutants of hypomorphic mutant allele for insulin-like growth factor 1 receptor (daf-2(e1370)) and a deletion mutation of the TOR-regulated S6K (rsks-1(ok1255)) synergistically increase lifespan compared to single mutants 20; their joint effect is greater than the sum of their individual effects. Analysis with a third genetic component revealed this two-factor interaction also required DAF-16, AMPK, HSF-1, and germline signaling. These gene interactions regulate lifespan via CYC-2 translation in the germline, which nonautonomously upregulates UPRmt and AMPK in the intestine to extend lifespan 21. Interactions between multiple components in nutrient sensing pathways are represented in Figure 3, and are detailed in Table 1.
Figure 3. Select Interactions between mechanisms/Hallmarks of aging.
Studies have investigated many interventions among multiple mechanisms or Hallmarks of aging, measuring outcomes upon higher levels in -omics and manifestation hierarchies. The manipulations include food supplementation, trans-genes, mutants, diet restriction, and drugs. We illustrate a select subset (not exhaustive) of interactions within and between mitochondrial, oxidative stress, nutrient sensing, and reproductive decline processes. Blue lines depict data for interactions between processes of two or more mechanisms. Green lines illustrate interactions within nutrient sensing for the widely studied subnetwork of insulin/IGF signaling.
Translation is also studied relative to nutrient sensing. Downregulated ribosomal subunits and elongation factors (RPS-15, RPS-22, S6K, eIF2β, or eIF4G) extend the lifespan of long-lived C. elegans eat-2 mutants, which have limited food intake by reducing pharyngeal pumping. Yet, downregulation of TOR/let-363 alone extends survival but does not increase eat-2 survival 22. Thus, different elements of translational control appear to be either dependent (let-363) or independent (elongation factors) of food intake (eat-2).
Direct manipulation of diet reveals how nutrients interact – or not – with regulatory mechanisms. C. elegans daf-2(e1370) long-lived mutants have been evaluated on five methods of diet manipulation 23. In each case, DR extended longevity independent of daf-2. Further studies manipulated the IIS transcription factor FOXO, encoded by daf-16 in C. elegans. Genetic interaction analysis has broadly established that daf-16 is required for the increased longevity of daf-2 mutants 24, although the interpretations do not account for the inherently shortened lifespan of the daf-16 mutant. A similar negative outcome was seen in Drosophila, where the longevity of dfoxo mutants was evaluated across a range of diet concentrations 25,26.
Ames dwarf mice can live 50% longer than their normal siblings. Ames have a defect in pituitary development (Prop1) that decreases circulating growth hormone (GH), prolactin (PRL), and thyrotropin (TSH) 27. The longevity of Ames mice appears independent of DR because restriction additively increases their longevity 28. DR, however, does not increase lifespan in long-lived growth hormone (GH) receptor knockout mice (GHRKO) 29. As in Drosophila, the effect of DR on mouse lifespan can be dissected from insulin signaling. Mice lacking adipose mTORC2 are insulin resistant, yet DR increases their lifespan as much as it does in wild-type mice 30. The impact of DR on lifespan also appears to be independent of Sirtuin3 (SIRT3), contrary to some expectations. Sirt3 knockout mice have wild-type lifespan, which can be increased by DR 31. Paradoxically, these interaction data collected across animal models provide little evidence to suggest that diet restriction modulates aging through insulin/IGF signaling.
Many studies evaluate whether drugs slow aging by inducing mechanisms of nutrient sensing 32,33. The TOR inhibitor rapamycin extends lifespan in yeast, worms, Drosophila and mice 34,35. However, rapamycin appears to not mediate longevity through IIS because the survival of daf-2 mutants treated with rapamycin is approximately additive to the impact of each intervention alone 20. Similarly, rapamycin extends the lifespan of long-lived Drosophila insulin receptor substrate chico heterozygote mutants 36. Although it is unknown whether rapamycin and insulin/IGF interact in mouse aging, these interventions have striking differences upon fat mass, insulin sensitivity 37, liver transcriptome and metabolome 38, and white adipose tissue 39. Once again and contrary to expectation, there is surprisingly scant evidence to support whether and how DR, IIS and TOR interact to modulate aging.
Interactions within the Mitochondrial Hallmark
Mitochondrial decline is a Hallmark of Aging associated with mtDNA mutations, elevated ROS and decreased oxidative phosphorylation (OXPHOS) 40. Knockdown of individual mitochondrial ETC and ATP synthase subunits extend C. elegans lifespan 41–43, but joint manipulations of these components have complex outcomes (Figure 2). Worm lifespan is increased by mutation of ISP-1 (isp-1(qm150)), a catalytic subunit of mitochondrial complex III 41. Similarly, lifespan is extended by mutation of NUO-6 (nuo-6(qm200)), a subunit of mitochondrial complex I (NDUFB4 ⁄ B15) 44. The lifespan of nuo-6, isp-1 double mutants was not different from that of single mutants, suggesting these subunits act within the same longevity pathway. However, survival induced by RNAi of either nuo-6 or isp-1 was additive to lifespan of isp-1(qm150, and nuo-6(qm200) mutants, suggesting these subunits act independently 44. The discrepancy among these inferences may arise because RNAi adds a temporal feature to the interaction, or because the partial penetrance afforded by RNAi uncovers unique, alternative aging-regulatory networks. This interaction demonstrates the importance of temporal regulation as well as the choice of genetic tools in order to detect an interaction between different components.
Figure 2. An example of interactions within a Hallmark: Mitochondria-related mechanisms.
Mitochondrial dynamics and function degrade with age: mitochondrial dysfunction is an identified Hallmark of Aging. Interactions between key aspects of mitochondrial function, including fission, fusion, oxidative phosphorylation, and translation have been studied in the context of aging. Green lines represent interactions identified to impact aging in at least one species.
Kayser et al. 45 demonstrated an interesting interaction between gas-1 and T26A5.3, orthologs of mitochondrial membrane respiratory chain NADH dehydrogenase (complex I) subunits. gas-1 and T26A5.3 mutant worms are individually short-lived, whereas T26A5.3; gas-1(fc21) double mutant worms are extremely long-lived, representing an interesting class of interactions in which the combination of two short-lived mutants results in an extended lifespan.
Fission and fusion dynamics control mitochondrial homeostasis 46. Mitochondrial fusion can be inhibited when eat-3 (worm orthologue of mammalian OPA-1) or fzo-1 (worm orthologue of human mitofusins MFN1 and MFN2) are depleted. Lifespan is synergistically increases when either of these interventions is combined with RNAi to inhibit mitochondrial translation (RNAi against mrp-5, mitochondrial ribosomal protein 5) 47. These interactions localize HLH-30 to the nucleus where the transcription factor induces lysosome biogenesis and autophagy. In this example, targeting different mitochondrial processes leads to a combined effect on lifespan.
Proteins involved in post-transcriptional modification of mitochondrial tRNAs also interact to control longevity 48. C. elegans MTCU-2 is the orthologue of human MTO1, and MTTU-1 is the orthologue of TRMU. These proteins respectively taurino-methylate and thiolate mt-tRNALys, mt-tRNAGlu, and mt-tRNAGln 49. Mutations of mttu-1 or mtcu-2 mildly reduce mitochondrial respiration. While mttu-1 mutants moderately increase lifespan, lifespan was increased more than 50% by the mtcu-2;mttu-1 double mutant. Interestingly, knockdown of daf-2 or rict-1 (a component of the target of rapamycin complex 2 (TORC2)) further extended the lifespan of the mtcu-2;mttu-1 double mutant. Impaired mitochondrial can further interact with IIS or TORC2 to modulate lifespan 48.
III. Interactions between different mechanisms/Hallmarks of aging
In the view of Geroscience, common mechanisms drive the progression of multiple Hallmarks of Aging 7,17. This hypothesis can be explored by interventions that affect the interaction of different hallmarks/mechanisms of aging (Figure 3). We illustrate this process through two combinations: Nutrient sensing with Reproduction, and Nutrient sensing with Redox. Other combinations are highlighted in Figure 4 and Supplementary Table 1.
Figure 4. Overview of interactions between different processes (hallmarks) of aging.
Various studies have combined approaches from different Hallmarks of Aging to test for additive or synergistic effects. This wheel, modified from the Hallmarks of Aging framework7,17, shows examples of various aging approaches that have been previously combined to further extend lifespan in three species: C. elegans, Drosophila, and mice. Blue lines represent combinatorial approaches that have shown promise in aging studies in at least one species. While most studies have been conducted in worms, targeting these combinations in multiple species can highlight promising approaches from a translational standpoint.
Nutrient sensing with Reproduction
Many studies show that lifespan is synergistically extended by the interaction of insulin and IGF-1 signaling (IIS) with reproduction. Lifespan is extended nearly four-fold in C. elegans when depletion of daf-2 is combined with germline knockout 50,51. This synergy suggests the longevity mechanisms underlying nutrient sensing and reproduction interact to amplify higher order manifestations of C. elegans aging. In contrast, Drosophila homozygous for null mutants of insulin receptor substrate homolog chico are sterile and increase median lifespan by about 50%, yet heterozygous chico females are equally as long-lived but are abundantly fertile 52,53. As well, inhibition of reproduction via the ovoD mutation 54 increases lifespan by ~14% 55 but flies heterozygous for both ovoD and chico only increase lifespan by 36%. In Drosophila it seems that nutrient sensing and reproduction act through common mechanisms to modulate lifespan. Results of interaction studies that are inconsistent between species may still illuminate some mechanisms, providing researchers with useful information about interactions in aging; however, these combinations may also have limited translatability across species.
Steroid hormones are key regulators of reproduction and aging. DAF-12 is the C. elegans ortholog of vertebrate vitamin D and liver X receptors. DAF-12 regulates developmental commitment to reproductive growth versus dauer, while in adults it mediates how signals from the gonad impact longevity 56. Although daf-12 mutation slightly shortens adult lifespan, lifespan is doubled when mutants of daf-12 are combined with a mutation of the insulin receptor 57. This synergy was completely suppressed by mutation of the foxo transcription factor daf-16, indicating that daf-16 acts downstream of both insulin and reproductive-steroid signaling.
Steroid sulfotransferases and sulfatases modulate sulfated hormones. Sulfatation impairs the ability of hormones to activate their receptors. In C. elegans, inhibition of steroid sulfatase by sul-2 mutation or pharmacological inhibition (using the specific STS inhibitor STX64) increased lifespan and ameliorated protein aggregation 58. These effects required the daf-16, daf-12, kri-1, tcer-1, and daf-36 factors, which function to extend lifespan through ablation of the germline. As well, sul-2 mutations further extended the lifespan of daf-2 mutant worms, and increased the lifespan of long-lived daf-10(m79) mutant worms, which are defective in sensory cilia formation 58.
Together these examples illustrate in C. elegans that reproduction and nutrient sensing have synergistic effects on lifespan.
Nutrient sensing with Redox regulation
Lifespan across various species can be extended by numerous antioxidant interventions based on genetics and pharmacology 59. A variety of natural compounds possess antioxidant properties and delay aging in model organisms, but these likewise affect multiple targets such that it is difficult to ascribe a specific mechanism. The topography of interactions is complicated by their pleiotropy of downstream targets and their potential cross-interactions.
One systematic example involves Brachionus manjavacas (Rotifera), which has a two-week long lifespan that otherwise displays typical patterns of animal aging 60. Among 20 single antioxidants tested, none alone extended lifespan. However, among 60 two-way combinations, seven produced a significant effect on rotifer lifespan: trolox and beta-carotene (10% lifespan increase), trolox and L-carnosine (13% lifespan increase); trolox and N-acetyl cysteine (16% lifespan increase); trolox and EUK-8 (14% lifespan increase); N-acetyl cysteine and quercetin (14.5% lifespan increase); EUK-8 and L-carnosine (10% lifespan increase); and EUK-8 and indole-3-propionic acid (14% lifespan increase). Interestingly, none of the 20 three- and four-way antioxidant combinations resulted in significant rotifer lifespan extension. This analysis represents one of the rare systematic studies evaluating all possible interactions between most available antioxidants. The potential topography and feedback among these interactions illustrates the complexity of understanding whether or how the added lifespan was caused specifically by antioxidant activity.
Antioxidant and nutrient sensing interventions have also been combined. Pyrroloquinoline quinone (PQQ) is a polyphenolic antioxidant that enhances resistance to oxidative stress and extends C. elegans lifespan 61,62. PQQ further extends lifespan of long-lived IIS pathway mutants (daf-2(e1370) and age-1(hx546), as well as that of eat-2 mutants), suggesting the mechanism for PQQ longevity differs from that of nutrient sensing. Another natural antioxidant, the root extract of Rhodiola rosea (RS), possesses antioxidant properties and has been shown to protect flies and human cells against oxidative stress, and to extend lifespan in yeast, worms, and flies 63–67. As with PQQ, RS extended lifespan of Drosophila with reduced nutrient sensing (wt/chico1) 67.
Superoxide dismutases (SODs) catalyze the conversion of superoxide into oxygen and hydrogen peroxide. In contrast to yeast, flies, and mice, where deletion of either cytoplasmic or mitochondrial SOD decreases lifespan; eliminating any of the five individual sod genes present in worms does not affect lifespan despite their increased sensitivity to oxidative stress. Instead, deletion of sod-2 extends lifespan even when sod-2 is deleted with other sod genes in double or triple sod mutants 68. Mutants of sod2 differentially interact with other Hallmarks: it is synergistic with the long-lived clk-1 mutant (Mitochondria), but only slightly increases the lifespans of long-lived eat-2 (Nutrient Sensing) and glp-1 (Reproduction) mutants. The observed synergy with clk suggests there are mechanistic interactions between redox signaling and mitochondrial function 68.
IV. High order interactions
Interaction analysis can simultaneously involve many targets (Table 2 and Table 3). Yen et al. manipulated four interventions known to extend C. elegans lifespan: low temperature, DR, reduced IlS, and mutations in mitochondrial genes 69. Double and triple interactions produced additive effects on lifespan, while DR and cold temperature were synergistic, implying that these two interventions extend lifespan via independent mechanisms 69. The strongest increase in the extension of lifespan was observed in the combined clk1/low temperature/DR or clk-1/daf-2/low temperature/DR interventions.
Table 2. Triple interventions for extension of lifespan.
Studies that have manipulated three nodes within one, two, or three Hallmarks of Aging and have shown an additive, partial, or synergistic effect on lifespan. Select studies of interest are summarized here, with the extent of effect added where relevant.
Organism | Intervention I | Signaling pathway (Hallmark of aging) I | Intervention II | Signaling pathway (Hallmark of aging) II | Intervention III | Signaling pathway (Hallmark of aging) III | % lifespan increase (combination) | Reference |
---|---|---|---|---|---|---|---|---|
C. elegans | mtcu-2 | Mitochondrial dysfunction | mttu-1 | Mitochondrial dysfunction | daf-2 | Deregulated nutrient sensing | ~70% | 48 |
C. elegans | mtcu-2 | Mitochondrial dysfunction | mttu-1 | Mitochondrial dysfunction | rict-1 | TORC2 | ~80% | 48 |
C. elegans | glp-1 | Reproduction | mrps-5 | Mitochondrial dysfunction | eat-3 | Mitochondrial dysfunction | 9 | |
C. elegans | ule-4 | Uterine lumen-expressed proteins | ule-5 | Uterine lumen-expressed proteins | far-6 | Uterine lumen-expressed proteins | 10–20% | 120 |
C.elegans | daf-2 | Deregulated nutrient sensing | clk-1 | Mitochondrial function | Low temperature | ~230% | 69 | |
C.elegans | DR | Deregulated nutrient sensing | clk-1 | Mitochondrial function | Low temperature | ~315% | 69 | |
C.elegans | daf-2 | Deregulated nutrient sensing | rsks-1 | Deregulated nutrient sensing | tax-6 | Calceneurine complex | 72 | |
C.elegans | eat-2 (ad1116) | Deregulated nutrient sensing | rifampicin | antibiotic | psora-4 | Kv1.3 channel blocker | 94 |
Table 3. Quadruple interventions for extension of lifespan.
Studies that have manipulated four nodes within multiple Hallmarks of Aging and have shown an additive, partial, or synergistic effect on lifespan. Select studies of interest are summarized here, with the extent of effect added where relevant.
Organism | Intervention I | Intervention II | Intervention III | Interventi on IV | Signaling pathways | |||
---|---|---|---|---|---|---|---|---|
C. elegans | ule-4 | ule-5 | ule-3 | ule-2 | Uterine lumen-expressed proteins | 15–20% | 120 | |
C. elegans | aakg-2(sta2) | ucp2 | lyz | hsf-1 | AMP activated protein kinase, mitochondrial uncoupling, heat shock transcription | ~130% | 70 | |
C. elegans | aakg-2(sta2) + lyz | ucp2 | unc-62 | hsf-1 | AMP activated protein kinase, mitochondrial uncoupling, heat shock transcription, neurotransmission neurotransmission | ~160% | 71 | |
D. melanogaster | Dark | 18 °C Dark | DR | 3G | maintaining in the dark, DR, low temperature, combination of rapamycin, berberine, and fucoxanthin | ~120% | 124 |
Aging interactions can also be probed using genes from multiple species at the same time 70. C. elegans lifespan is extended by overexpressing hsf-1 (heat shock transcription factor), aakg-2 (gamma subunit of AMP-activated protein kinase), sod-1 (cytosolic superoxide dismutase), and Imp-2 (chaperone-mediated autophagy). Introduced genes from zebrafish likewise extended lifespan, including ucp2 (mitochondrial uncoupling) and lyz (anti-bacterial enzyme). Combining aakg-2(sta2) and ucp2 extended lifespan ~80%, while hsf-1 and lyz extended lifespan ~60%. Combinations of three genes extended lifespan about 100% (aakg-2, ucp2, lyz or aakg-2, hsf-1, lyz). Combining all four genes together impressively extended lifespan ~135%, and these worms were more resistant to paraquat than those expressing two or three genes together 70. Downregulation of HOX co-factor unc-62 (Homothorax) (quintuply-modified C. elegans strain) further extended lifespan by 15% relative to the quadruple strain and by 160% relative to the control strain 71. Downregulation of unc-62 exerted several beneficial effects on lifespan. It prevented the accumulation of yolk proteins in old worms, the age-dependent overall decline of transcription in the intestine, and the age-dependent upregulated expression of neuronal genes in the intestine 71.
While dietary restriction has surprisingly weak interactions in two-way analyses, it produces connections when studied with higher order approaches. Three regulatory modules were identified in C. elegans undergoing calorie restriction or intermittent fasting: rheb-1_let-363/tor, aak-2_tax-6_xbp-1, and daf-16_glp-1 72. Downregulating all three modules simultaneously additively extended lifespan (175%) and the lifespan of these worms remained insensitive to DR 72.
Higher order combinations can also be studied in mammals. Davidsohn et al. queried fibroblast growth factor 21 (FGF21), αKlotho, and transforming growth factor-β1 (TGFβ1) 73, which are secreted factors with beneficial roles in aging 74–76. Each was tested alone and together through adeno-associated virus (AAV)-mediated gene therapy for their impact on age-related diseases: obesity, type II diabetes, heart failure, and renal failure. FGF21 alone reversed obesity and diabetic phenotypes. The combination of AAV:sTGFβR2 combined with FGF21 prevented renal medullary atrophy (unilateral ureteral obstruction was used as a model of progressive renal fibrosis, which is a feature of many forms of progressive renal diseases). Ascending aortic constriction was used to simulate heart failure, mimicking age-related hypertrophy caused by systemic hypertension. TGFβR2 with either FGF21 or αKlotho moderated heart failure. Overall, TGFβR2 with FGF21 successfully treated all four age-related diseases at once 73.
V. Pharmacological interactions
Beyond how interacting mechanisms coordinate through the Hallmarks of Aging, pharmacological interaction analysis empirically addresses how simultaneous interventions maximize successful aging (Table 4). One approach uses drugs already recognized for the effects on conditions with known age-associated progression, such as diabetes or cardiovascular disease. A broader strategy uses unbiased or targeted pharmacological screens based on model organism lifespan. In both cases, drug testing adds the ability to explore treatments limited to adults, or even at post-reproductive stages. On the other hand, compounds often target multiple pathways, and here it will be important to evaluate the combined drug treatments with specific genetic mutants in a higher order analysis.
Table 4. Pharmacological interventions for extension of lifespan.
Studies that have explored pharmacological interventions, particularly those that are already approved for use in humans, are appealing due to their translational potential. Several promising candidate combinations are summarized here, with the extent of effect added where relevant.
Organism | Intervention I | Signaling pathway (Hallmark of aging) I | Intervention II | Signaling pathway (Hallmark of aging) II | Intervention III | Signaling pathway (Hallmark of aging) III | % lifespan increase (combination) | Reference |
---|---|---|---|---|---|---|---|---|
C. elegans | Valproic acid | antiepileptic | trimeth adione (TRI) | anticonvulsant | 63% | 93 | ||
C. elegans | Valproic acid | antiepileptic | kanamycin | antibiotic | 41% | 93 | ||
C. elegans | Allantoin | rifampicin | antibiotic | rapamycin | mTO RC1 inhibitor | 89% | 94 | |
C. elegans | Allantoin | rifampicin | antibiotic | psora-4 | Kv1. 3 channel blocker | 96% | 94 | |
D. melanogaster | trametinib | MEK inhibitor | rapamycin | mTORC1 inhibitor | lithium | GSK-3 inhibitor | 48% | 97 |
s. cerevisiae | rapamycin | mTORC1 inhibitor | myriocin | sphingolipid synthesis inhibitor | 102 | |||
S. pombe | rapamycin | mTORCI inhibitor | myriocin | sphingolipid synthesis inhibitor | 103 |
Dasatinib and quercetin
Based on a wealth of single intervention treatments, eliminating senescent cells extends mouse healthspan 77. The combination of dasatinib (inhibitor of multiple tyrosine kinases) and quercetin (a natural flavonol), both approved for use in humans, was more effective in eliminating senescent cells than either of these drugs alone 78. Likewise, treating a Ercc1−/Δ mouse model of accelerated aging with dasatinib and quercetin weekly significantly reduced senescence markers and frailty, and extended healthspan 78. Biweekly administration of dasatinib and quercetin to wild-type mice starting at 24–27 months increased posttreatment lifespan by 36%, lowered mortality hazard, and improved healthspan 79. Currently, clinical trials in humans are testing how various age-related diseases respond to dasatinib combined with quercetin (ClinicalTrials.gov).
Rapamycin and metformin
Metformin is a widely prescribed FDA-approved oral antidiabetic drug that may target several molecular mechanisms associated with aging 80. To date there are mixed results on whether metformin can slow aging based on rodent models. It modestly extends the lifespan of tumor-prone HER2/neu mice by 8% 81, and of male C57BL/6 mice when begun at middle age 82. However, in the NIA Interventions Testing Program, treating mice with 0.1% metformin in the diet alone did not significantly extend lifespan 83. Yet, because metformin has protective age-associated effects in diabetics, trials are envisioned to measure its potential to delay normal human aging 84. In this context, dual interventions may be promising. Notably, metformin combined with rapamycin robustly extended mouse lifespan compared to an independently tested cohort treated with rapamycin alone 83. However, it is unclear whether the lifespan with rapamycin and metformin would be extended compared to the single treatments in a properly controlled experiment.
Rapamycin and acarbose
Acarbose is an FDA-approved anti-diabetic drug. In UM-HET3 mice it increased male longevity, decreased liver degeneration in males and females, decreased lung tumors in males, and improved rotarod performance in aged females 85. The combination of rapamycin and acarbose administered to male mice starting at 9 months of age resulted in a significant increase in lifespan across three separate study sites, and this increase was higher than historical data for treatment with rapamycin alone 86. A further combination of rapamycin, acarbose, and phenylbutyrate improved age-related phenotypes including body composition, cognition, strength, endurance, and tissue resilience 87. As expected, rapamycin can have negative side effects including glucose intolerance and insulin resistance 88,89. The antidiabetic drugs metformin and acarbose may alleviate these detrimental effects 90, and thereby enhance how rapamycin impacts lifespan.
Simvastatin and ramipril
Simvastatin and ramipril are routinely used to treat cardiovascular disease (CVD) 91. Simvastatin reduces the biosynthesis of the isoprenoids used for cholesterol biosynthesis, while ramipril reduces the biosynthesis of angiotensin and the activity of the angiotensin receptors. Long-lived, B6C3F1 mice fed simvastatin along with ramipril increased mouse lifespan by 9%, while simvastatin or ramipril alone were ineffective. The combination also promoted weight loss despite eating the same number of calories as controls 91.
Valproic acid and trimethadione
Several anticonvulsants (ethosuximide, trimethadione, and 3,3-diethyl-2-pyrrolidinone/DEABL) extend worm healthspan and lifespan 92. Longevity conferred by ethosuximide and trimethadione appears to be independent of DR. Similarly, trimethadione additively extended lifespans of long-lived worms carrying loss of-function mutations of genes important for the function of sensory neurons (osm-3 and tax-4), for neurotransmission (unc-31, unc-64, and aex-3), as well as daf-2 (insulin/IGF-1 receptor) mutant worms 92. The antiepileptic drug Valproic Acid (VA) was also demonstrated to extend worm lifespan up to 35%, as does structurally similar Valpromide (VPD). While Ethosuximide and trimethadione (TRI) extended lifespan of daf-16 mutant worms, VA does not 93. This is consistent with how worms treated with a combination of VA and TRI lived longer than mono-treatment, suggesting these drugs act via separate mechanisms 93. In addition, although VA might inhibit bacterial proliferation and extend worm lifespan via inhibition of bacterial pathogenicity, the combination of VA and kanamycin had a partially additive effect on extension of lifespan, implying a mechanism independent of bacterial proliferation inhibition 93.
Allantoin, rifampicin, rapamycin, and psora-4
Admasu et al. defined drug combinations by initially testing five drugs individually that were previously shown to extend C. elegans lifespan: rapamycin, rifampicin, metformin, psora-4, allantoin, 94. Using transcriptional profiling and principal component analysis, they found that rifampicin, rapamycin, and psora-4 cause the most distinctive effects. While each extends lifespan individually, rifampicin with rapamycin, and rifampicin with psora-4 increased survival more than the largest effect of any drug alone. When allantoin was added to these combinations, worm lifespan was further extended by about 90% 94. Rifampicin with psora-4 also extended the lifespan of eat-2(ad1116) worms, implying this combination acts independently of DR. Each synergistic drug combination likewise improved movement of aged worms, and had higher resistance to thermal and oxidative stress (paraquat). Similar to worms, the longevity of male Drosophila was synergistically increased by the combinations of rapamycin with rifampicin, and rapamycin with rifampicin and allantoin 94.
Trametinib (MEK inhibitor), rapamycin (mTORC1 inhibitor), and lithium (GSK-3 inhibitor)
Trametinib is a MEK inhibitor used to manage cancer, while lithium is a GSK-3 inhibitor used to treat mood disorders. Trametinib and lithium each increase Drosophila lifespan alone and further extends the lifespan of flies lacking the insulin-like peptides 2, 3, and 5 (dilp2–3,5) 95,96. Double combinations of lithium and rapamycin, lithium and trametinib, or rapamycin and trametinib produced greater lifespan extension than any single interventions, and longevity was increased nearly 50% when all three drugs were combined 97.
Rapamycin and wortmannin
Danilov et al. tested rapamycin and other anticancer agents for their ability to extend Drosophila health- and lifespan. They tested the effects of rapamycin (mTORC1 inhibitor), wortmannin (PI3K inhibitor), 1400W (iNOS inhibitor), pyrrolidin dithiocarbamate (NF-kB inhibitor), and QNZ (NF-kB inhibitor) as single agents or in combination. The most effective lifespan-expanding intervention was the combination of rapamycin and wortmannin, which extended Drosophila lifespan by 23.4% 98.
Rapamycin and myriocin
Lipid metabolism plays a key role in the regulation of aging and longevity 99–101. Myriocin is a potent inhibitor of serine palmitoyltransferase, the first step in sphingosine biosynthesis. When combined with a low dose of rapamycin, chronological lifespan was synergistically increased in both S. cerevisiae and S. pombe 102 103. This dual treatment likewise increased resistance to heat and oxidative stress, and increased genomic stability 102.
Metformin with SRT1720 reduce lifespan
Paradoxically, molecules that individually extend healthspan and lifespan may reduce survival when combined. Two SIRT1 activators (SRT2104 and SRT1720) improve the survival and age-associated health of mice on high fat diet that induces traits of type 2 diabetes 104–107. Yet, such mice treated with both metformin and SRT1720 dramatically lose body weight and experience reduced survival. On the other hand, these mice were leaner and performed significantly better on the accelerating rotarod relative to age-matched controls 108.
VII. Conclusion and open questions
Dozens of genetic and pharmacological interventions interact to affect Hallmarks of Aging and lifespan. We need to address outstanding questions to synthesize these nodes into systems and networks.
How do we define the topology of multiple and varied interactions?
We started our review by discussing different ways to structure the complexity of aging across levels of interaction. “Hallmarks of Aging” 7,17 features as a common level for many studies, although aging research has yet to fix what qualifies or not as a Hallmark. We still require a clear classification of aging-related processes or how to hierarchically structure their interactions. We also require better concepts of feedback, both positive and negative, within interaction networks.
How do we predict which combined interventions will optimally slow aging?
To predict prospective cooperation between different targets, we seek similarity in downstream effectors of lifespan. For example, different lifespan-extending interventions in mice that inhibited protein translation 109, and the size of nucleoli 110 and flux of 1C metabolism 111 were affected by manipulations of different pro-longevity pathways in worms. Our challenge is to a priori select combinations of interventions that will synergistically amplify their individual effects upon these common targets, and thereby efficiently manage aging.
What new genetic tools are needed to test high-order combinations?
Most interactions to date have been described using C. elegans because strains are coisogenic and bacterial dsRNA offers easy combinatorial gene knockdown. To study multi-feature interactions in Drosophila or mice requires great care to manage genetic background and tools to overcome the constraints of transgenes. One solution might multiplex several RNAi 112 or gRNA lines 113 into one cassette with an easily identifiable genetic marker, as we previously applied to target several cancer-related genes simultaneously 114.
How do we sample or create a matrix of all pairwise interactions?
Norris et al. developed the CRISPR/Cas9-based Synthetic Genetic Interaction (CRISPR-SGI) approach to systematically generate double-mutants in C. elegans 115. They generated all possible single and double mutations of 14 conserved RNA-binding proteins and measured the effect of each on worm fitness. A similar approach can be applied to a large set of known longevity genes to comprehensively study how they interact to affect aging.
What constrains the potential to further slow aging?
Geroscience proposes that common underlying processes modulate the progression of multiple manifestations or Hallmarks of aging. If targeting two independent underlying processes additively affects aging, what determines when further interactions cease to increase health and lifespan? Targeting all defined hallmarks is unlikely to extend lifespan without limits. Are there “weak” spots that constrain further lifespan? Might these involve trade-offs between biological necessities such as reproduction and cellular integrity? Will multi-interventions produce side effects? As we elaborate the processes driving diverse Hallmarks of aging, can we understand when they become jointly constrained?
How do we predict which combined interventions apply to humans?
One approach might use epigenetic-like clocks, which predict chronological and biological age in humans 116. One application of such tools is to predict how interventions at younger ages will affect the trajectory of aging. This approach may be extended to how multiple interventions interact to affect aging. For instance, the Thymus Regeneration, Immunorestoration, and Insulin Mitigation clinical trial found epigenetic age was reduced by approximately 1.5 years after one year of treatment with a combination of human growth hormone (rhGH), dehydroepiandrosterone (DHEA), and metformin 117. As interaction analyses expand from model systems to applications for humans we will need robust approaches to evaluate effectiveness and safety, and all within constrained timeframes.
Supplementary Material
Acknowledgments:
This work was supported by NIGMS R35 GM146869 (P.A.), NIA R00 AG057792 (P.A.), NIA R03 AG075651 (P.A.), NIA P30 AG024827 pilot grant (P.A.), Richard King Mellon Foundation award (P.A.), NIA R01 AG059563 (T.M.), and NIA R01 AG069639 (T.M.).
Footnotes
Competing interests
The authors declare that they have no competing interests.
Editor summary:
Parthitko et al discuss combinatorial approaches targeting underlying mechanisms of aging across species and describe frameworks to analyze these interactions and their cross-species translational potential.
References:
- 1.de Magalhaes JP & Toussaint O GenAge: a genomic and proteomic network map of human ageing. FEBS letters 571, 243–247, doi: 10.1016/j.febslet.2004.07.006 (2004). [DOI] [PubMed] [Google Scholar]
- 2.Barardo DG et al. Machine learning for predicting lifespan-extending chemical compounds. Aging (Albany NY) 9, 1721–1737, doi: 10.18632/aging.101264 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Parkhitko AA, Filine E, Mohr SE, Moskalev A & Perrimon N Targeting metabolic pathways for extension of lifespan and healthspan across multiple species. Ageing research reviews 64, 101188, doi: 10.1016/j.arr.2020.101188 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Moskalev A et al. Targeting aging mechanisms: pharmacological perspectives. Trends Endocrinol Metab 33, 266–280, doi: 10.1016/j.tem.2022.01.007 (2022). [DOI] [PubMed] [Google Scholar]
- 5.Hoffman JM et al. Effects of age, sex, and genotype on high-sensitivity metabolomic profiles in the fruit fly, Drosophila melanogaster. Aging cell 13, 596–604, doi: 10.1111/acel.12215 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Parkhitko AA, Jouandin P, Mohr SE & Perrimon N Methionine metabolism and methyltransferases in the regulation of aging and lifespan extension across species. Aging cell 18, e13034, doi: 10.1111/acel.13034 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Lopez-Otin C, Blasco MA, Partridge L, Serrano M & Kroemer G Hallmarks of aging: An expanding universe. Cell 186, 243–278, doi: 10.1016/j.cell.2022.11.001 (2023). [DOI] [PubMed] [Google Scholar]
- 8.Golubev AG An essay on the nominal vs. real definitions of aging. Biogerontology 22, 441–457, doi: 10.1007/s10522-021-09926-x (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Liu YJ, McIntyre RL, Janssens GE & Houtkooper RH Mitochondrial fission and fusion: A dynamic role in aging and potential target for age-related disease. Mechanisms of ageing and development 186, 111212, doi: 10.1016/j.mad.2020.111212 (2020). [DOI] [PubMed] [Google Scholar]
- 10.Riuzzi F et al. Cellular and molecular mechanisms of sarcopenia: the S100B perspective. J Cachexia Sarcopenia Muscle 9, 1255–1268, doi: 10.1002/jcsm.12363 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.He Y et al. Cellular Senescence in Sarcopenia: Possible Mechanisms and Therapeutic Potential. Front Cell Dev Biol 9, 793088, doi: 10.3389/fcell.2021.793088 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Phillips PC The language of gene interaction. Genetics 149, 1167–1171, doi: 10.1093/genetics/149.3.1167 (1998). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Gems D, Pletcher S & Partridge L Interpreting interactions between treatments that slow aging. Aging cell 1, 1–9 (2002). [DOI] [PubMed] [Google Scholar]
- 14.Kang J et al. Ribosomal proteins and human diseases: molecular mechanisms and targeted therapy. Signal Transduct Target Ther 6, 323, doi: 10.1038/s41392-021-00728-8 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Ferre M, Amati-Bonneau P, Tourmen Y, Malthiery Y & Reynier P eOPA1: an online database for OPA1 mutations. Human mutation 25, 423–428, doi: 10.1002/humu.20161 (2005). [DOI] [PubMed] [Google Scholar]
- 16.Berdynski M et al. SOD1 mutations associated with amyotrophic lateral sclerosis analysis of variant severity. Sci Rep 12, 103, doi: 10.1038/s41598-021-03891-8 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Lopez-Otin C, Blasco MA, Partridge L, Serrano M & Kroemer G The hallmarks of aging. Cell 153, 1194–1217, doi: 10.1016/j.cell.2013.05.039 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Mannick JB & Lamming DW Targeting the biology of aging with mTOR inhibitors. Nat Aging, doi: 10.1038/s43587-023-00416-y (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Tatar M, Bartke A & Antebi A The endocrine regulation of aging by insulin-like signals. Science (New York, N.Y 299, 1346–1351, doi: 10.1126/science.1081447 (2003). [DOI] [PubMed] [Google Scholar]
- 20.Chen D et al. Germline signaling mediates the synergistically prolonged longevity produced by double mutations in daf-2 and rsks-1 in C. elegans. Cell reports 5, 1600–1610, doi: 10.1016/j.celrep.2013.11.018 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Lan J et al. Translational Regulation of Non-autonomous Mitochondrial Stress Response Promotes Longevity. Cell reports 28, 1050–1062 e1056, doi: 10.1016/j.celrep.2019.06.078 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Hansen M et al. Lifespan extension by conditions that inhibit translation in Caenorhabditis elegans. Aging cell 6, 95–110, doi: 10.1111/j.1474-9726.2006.00267.x (2007). [DOI] [PubMed] [Google Scholar]
- 23.Greer EL & Brunet A Different dietary restriction regimens extend lifespan by both independent and overlapping genetic pathways in C. elegans. Aging Cell 8, 113–127 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Kenyon C, Chang J, Gensch E, Rudner A & Tabtiang R A C. elegans mutant that lives twice as long as wild type. Nature 366, 461–464, doi: 10.1038/366461a0 (1993). [DOI] [PubMed] [Google Scholar]
- 25.Giannakou ME, Goss M & Partridge L Role of dFOXO in lifespan extension by dietary restriction in Drosophila melanogaster: not required, but its activity modulates the response. Aging cell 7, 187–198, doi: 10.1111/j.1474-9726.2007.00362.x (2008). [DOI] [PubMed] [Google Scholar]
- 26.Min KJ, Yamamoto R, Buch S, Pankratz M & Tatar M Drosophila lifespan control by dietary restriction independent of insulin-like signaling. Aging cell 7, 199–206, doi: 10.1111/j.1474-9726.2008.00373.x (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Brown-Borg HM, Borg KE, Meliska CJ & Bartke A Dwarf mice and the ageing process. Nature 384, 33, doi: 10.1038/384033a0 (1996). [DOI] [PubMed] [Google Scholar]
- 28.Bartke A et al. Extending the lifespan of long-lived mice. Nature 414, 412, doi: 10.1038/35106646 (2001). [DOI] [PubMed] [Google Scholar]
- 29.Bonkowski MS, Rocha JS, Masternak MM, Al Regaiey KA & Bartke A Targeted disruption of growth hormone receptor interferes with the beneficial actions of calorie restriction. Proceedings of the National Academy of Sciences of the United States of America 103, 7901–7905, doi: 10.1073/pnas.0600161103 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Yu D et al. Calorie-Restriction-Induced Insulin Sensitivity Is Mediated by Adipose mTORC2 and Not Required for Lifespan Extension. Cell reports 29, 236–248 e233, doi: 10.1016/j.celrep.2019.08.084 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Dhillon RS et al. SIRT3 deficiency decreases oxidative metabolism capacity but increases lifespan in male mice under caloric restriction. Aging cell 21, e13721, doi: 10.1111/acel.13721 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Hofer SJ, Davinelli S, Bergmann M, Scapagnini G & Madeo F Caloric Restriction Mimetics in Nutrition and Clinical Trials. Front Nutr 8, 717343, doi: 10.3389/fnut.2021.717343 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Madeo F, Pietrocola F, Eisenberg T & Kroemer G Caloric restriction mimetics: towards a molecular definition. Nat Rev Drug Discov 13, 727–740, doi: 10.1038/nrd4391 (2014). [DOI] [PubMed] [Google Scholar]
- 34.Parkhitko AA, Favorova OO, Khabibullin DI, Anisimov VN & Henske EP Kinase mTOR: regulation and role in maintenance of cellular homeostasis, tumor development, and aging. Biochemistry. Biokhimiia 79, 88–101, doi: 10.1134/S0006297914020023 (2014). [DOI] [PubMed] [Google Scholar]
- 35.Selvarani R, Mohammed S & Richardson A Effect of rapamycin on aging and age-related diseases-past and future. Geroscience 43, 1135–1158, doi: 10.1007/s11357-020-00274-1 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Bjedov I et al. Mechanisms of life span extension by rapamycin in the fruit fly Drosophila melanogaster. Cell metabolism 11, 35–46, doi: 10.1016/j.cmet.2009.11.010 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Fok WC et al. Short-term treatment with rapamycin and dietary restriction have overlapping and distinctive effects in young mice. The journals of gerontology. Series A, Biological sciences and medical sciences 68, 108–116, doi: 10.1093/gerona/gls127 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Fok WC et al. Combined treatment of rapamycin and dietary restriction has a larger effect on the transcriptome and metabolome of liver. Aging cell 13, 311–319, doi: 10.1111/acel.12175 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Fok WC et al. Short-term rapamycin treatment in mice has few effects on the transcriptome of white adipose tissue compared to dietary restriction. Mechanisms of ageing and development 140, 23–29, doi: 10.1016/j.mad.2014.07.004 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Bratic A & Larsson NG The role of mitochondria in aging. The Journal of clinical investigation 123, 951–957, doi: 10.1172/JCI64125 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Feng J, Bussiere F & Hekimi S Mitochondrial electron transport is a key determinant of life span in Caenorhabditis elegans. Developmental cell 1, 633–644 (2001). [DOI] [PubMed] [Google Scholar]
- 42.Dillin A et al. Rates of behavior and aging specified by mitochondrial function during development. Science (New York, N.Y 298, 2398–2401, doi: 10.1126/science.1077780 (2002). [DOI] [PubMed] [Google Scholar]
- 43.Copeland JM et al. Extension of Drosophila life span by RNAi of the mitochondrial respiratory chain. Curr Biol 19, 1591–1598, doi: 10.1016/j.cub.2009.08.016 (2009). [DOI] [PubMed] [Google Scholar]
- 44.Yang W & Hekimi S Two modes of mitochondrial dysfunction lead independently to lifespan extension in Caenorhabditis elegans. Aging cell 9, 433–447, doi: 10.1111/j.1474-9726.2010.00571.x (2010). [DOI] [PubMed] [Google Scholar]
- 45.Kayser EB, Sedensky MM & Morgan PG The effects of complex I function and oxidative damage on lifespan and anesthetic sensitivity in Caenorhabditis elegans. Mechanisms of ageing and development 125, 455–464, doi: 10.1016/j.mad.2004.04.002 (2004). [DOI] [PubMed] [Google Scholar]
- 46.Youle RJ & van der Bliek AM Mitochondrial fission, fusion, and stress. Science (New York, N.Y 337, 1062–1065, doi:337/6098/1062 [pii] 10.1126/science.1219855 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Liu YJ et al. Mitochondrial translation and dynamics synergistically extend lifespan in C. elegans through HLH-30. The Journal of cell biology 219, doi: 10.1083/jcb.201907067 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Navarro-Gonzalez C et al. Mutations in the Caenorhabditis elegans orthologs of human genes required for mitochondrial tRNA modification cause similar electron transport chain defects but different nuclear responses. PLoS Genet 13, e1006921, doi: 10.1371/journal.pgen.1006921 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Suzuki T & Suzuki T A complete landscape of post-transcriptional modifications in mammalian mitochondrial tRNAs. Nucleic acids research 42, 7346–7357, doi: 10.1093/nar/gku390 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Arantes-Oliveira N, Berman JR & Kenyon C Healthy animals with extreme longevity. Science (New York, N.Y 302, 611, doi: 10.1126/science.1089169 (2003). [DOI] [PubMed] [Google Scholar]
- 51.Hsin H & Kenyon C Signals from the reproductive system regulate the lifespan of C. elegans. Nature 399, 362–366, doi: 10.1038/20694 (1999). [DOI] [PubMed] [Google Scholar]
- 52.Tu MP, Epstein D & Tatar M The demography of slow aging in male and female Drosophila mutant for the insulin-receptor substrate homologue chico. Aging cell 1, 75–80, doi: 10.1046/j.1474-9728.2002.00010.x (2002). [DOI] [PubMed] [Google Scholar]
- 53.Clancy DJ et al. Extension of life-span by loss of CHICO, a Drosophila insulin receptor substrate protein. Science (New York, N.Y 292, 104–106, doi: 10.1126/science.1057991 (2001). [DOI] [PubMed] [Google Scholar]
- 54.Oliver B, Perrimon N & Mahowald AP The ovo locus is required for sex-specific germ line maintenance in Drosophila. Genes & development 1, 913–923, doi: 10.1101/gad.1.9.913 (1987). [DOI] [PubMed] [Google Scholar]
- 55.Sgro CM & Partridge L A delayed wave of death from reproduction in Drosophila. Science (New York, N.Y 286, 2521–2524 (1999). [DOI] [PubMed] [Google Scholar]
- 56.Antebi A Steroid regulation of C. elegans diapause, developmental timing, and longevity. Curr Top Dev Biol 105, 181–212, doi: 10.1016/B978-0-12-396968-2.00007-5 (2013). [DOI] [PubMed] [Google Scholar]
- 57.Larsen PL, Albert PS & Riddle DL Genes that regulate both development and longevity in Caenorhabditis elegans. Genetics 139, 1567–1583 (1995). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Perez-Jimenez MM et al. Steroid hormones sulfatase inactivation extends lifespan and ameliorates age-related diseases. Nature communications 12, 49, doi: 10.1038/s41467-020-20269-y (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Kregel KC & Zhang HJ An integrated view of oxidative stress in aging: basic mechanisms, functional effects, and pathological considerations. Am J Physiol Regul Integr Comp Physiol 292, R18–36, doi: 10.1152/ajpregu.00327.2006 (2007). [DOI] [PubMed] [Google Scholar]
- 60.Snell TW, Fields AM & Johnston RK Antioxidants can extend lifespan of Brachionus manjavacas (Rotifera), but only in a few combinations. Biogerontology 13, 261–275, doi: 10.1007/s10522-012-9371-x (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Wu JZ et al. Pyrroloquinoline quinone enhances the resistance to oxidative stress and extends lifespan upon DAF-16 and SKN-1 activities in C. elegans. Experimental gerontology 80, 43–50, doi: 10.1016/j.exger.2016.04.008 (2016). [DOI] [PubMed] [Google Scholar]
- 62.Sasakura H et al. Lifespan extension by peroxidase and dual oxidase-mediated ROS signaling through pyrroloquinoline quinone in C. elegans. Journal of cell science 130, 2631–2643, doi: 10.1242/jcs.202119 (2017). [DOI] [PubMed] [Google Scholar]
- 63.Schriner SE et al. Decreased mitochondrial superoxide levels and enhanced protection against paraquat in Drosophila melanogaster supplemented with Rhodiola rosea. Free Radic Res 43, 836–843, doi: 10.1080/10715760903089724 (2009). [DOI] [PubMed] [Google Scholar]
- 64.Schriner SE, Avanesian A, Liu Y, Luesch H & Jafari M Protection of human cultured cells against oxidative stress by Rhodiola rosea without activation of antioxidant defenses. Free Radic Biol Med 47, 577–584, doi: 10.1016/j.freeradbiomed.2009.05.025 (2009). [DOI] [PubMed] [Google Scholar]
- 65.Bayliak MM & Lushchak VI The golden root, Rhodiola rosea, prolongs lifespan but decreases oxidative stress resistance in yeast Saccharomyces cerevisiae. Phytomedicine 18, 1262–1268, doi: 10.1016/j.phymed.2011.06.010 (2011). [DOI] [PubMed] [Google Scholar]
- 66.Wiegant FA et al. Plant adaptogens increase lifespan and stress resistance in C. elegans. Biogerontology 10, 27–42, doi: 10.1007/s10522-008-9151-9 (2009). [DOI] [PubMed] [Google Scholar]
- 67.Schriner SE et al. Extension of Drosophila lifespan by Rhodiola rosea through a mechanism independent from dietary restriction. PLoS One 8, e63886, doi: 10.1371/journal.pone.0063886 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Van Raamsdonk JM & Hekimi S Deletion of the mitochondrial superoxide dismutase sod-2 extends lifespan in Caenorhabditis elegans. PLoS Genet 5, e1000361, doi: 10.1371/journal.pgen.1000361 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Yen K & Mobbs CV Evidence for only two independent pathways for decreasing senescence in Caenorhabditis elegans. Age (Dordr) 32, 39–49, doi: 10.1007/s11357-009-9110-7 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Sagi D & Kim SK An engineering approach to extending lifespan in C. elegans. PLoS Genet 8, e1002780, doi: 10.1371/journal.pgen.1002780 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Sagi D The addition of a developmental factor, unc-62, to already long-lived worms increases lifespan and healthspan. Biol Open 6, 1796–1801, doi: 10.1242/bio.027433 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Hou L et al. A Systems Approach to Reverse Engineer Lifespan Extension by Dietary Restriction. Cell metabolism 23, 529–540, doi: 10.1016/j.cmet.2016.02.002 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Davidsohn N et al. A single combination gene therapy treats multiple age-related diseases. Proceedings of the National Academy of Sciences of the United States of America 116, 23505–23511, doi: 10.1073/pnas.1910073116 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Zhang Y et al. The starvation hormone, fibroblast growth factor-21, extends lifespan in mice. eLife 1, e00065, doi: 10.7554/eLife.00065 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Kurosu H et al. Suppression of aging in mice by the hormone Klotho. Science (New York, N.Y 309, 1829–1833, doi: 10.1126/science.1112766 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Brooks WW & Conrad CH Myocardial fibrosis in transforming growth factor beta(1)heterozygous mice. J Mol Cell Cardiol 32, 187–195, doi: 10.1006/jmcc.1999.1065 (2000). [DOI] [PubMed] [Google Scholar]
- 77.Baker DJ et al. Clearance of p16Ink4a-positive senescent cells delays ageing-associated disorders. Nature 479, 232–236, doi: 10.1038/nature10600 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Zhu Y et al. The Achilles' heel of senescent cells: from transcriptome to senolytic drugs. Aging cell 14, 644–658, doi: 10.1111/acel.12344 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Xu M et al. Senolytics improve physical function and increase lifespan in old age. Nat Med 24, 1246–1256, doi: 10.1038/s41591-018-0092-9 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Kulkarni AS, Gubbi S & Barzilai N Benefits of Metformin in Attenuating the Hallmarks of Aging. Cell metabolism, doi: 10.1016/j.cmet.2020.04.001 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Anisimov VN et al. Effect of metformin on life span and on the development of spontaneous mammary tumors in HER-2/neu transgenic mice. Experimental gerontology 40, 685–693, doi: 10.1016/j.exger.2005.07.007 (2005). [DOI] [PubMed] [Google Scholar]
- 82.Martin-Montalvo A et al. Metformin improves healthspan and lifespan in mice. Nature communications 4, 2192, doi: 10.1038/ncomms3192 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Strong R et al. Longer lifespan in male mice treated with a weakly estrogenic agonist, an antioxidant, an alpha-glucosidase inhibitor or a Nrf2-inducer. Aging cell 15, 872–884, doi: 10.1111/acel.12496 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Barzilai N, Crandall JP, Kritchevsky SB & Espeland MA Metformin as a Tool to Target Aging. Cell metabolism 23, 1060–1065, doi: 10.1016/j.cmet.2016.05.011 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Harrison DE et al. Acarbose improves health and lifespan in aging HET3 mice. Aging cell 18, e12898, doi: 10.1111/acel.12898 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Strong R et al. Lifespan benefits for the combination of rapamycin plus acarbose and for captopril in genetically heterogeneous mice. Aging cell 21, e13724, doi: 10.1111/acel.13724 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Jiang Z et al. Short term treatment with a cocktail of rapamycin, acarbose and phenylbutyrate delays aging phenotypes in mice. Sci Rep 12, 7300, doi: 10.1038/s41598-022-11229-1 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Lamming DW et al. Rapamycin-induced insulin resistance is mediated by mTORC2 loss and uncoupled from longevity. Science (New York, N.Y 335, 1638–1643, doi: 10.1126/science.1215135 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Lamming DW et al. Young and old genetically heterogeneous HET3 mice on a rapamycin diet are glucose intolerant but insulin sensitive. Aging cell 12, 712–718, doi: 10.1111/acel.12097 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Weiss R, Fernandez E, Liu Y, Strong R & Salmon AB Metformin reduces glucose intolerance caused by rapamycin treatment in genetically heterogeneous female mice. Aging (Albany NY) 10, 386–401, doi: 10.18632/aging.101401 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Spindler SR, Mote PL & Flegal JM Combined statin and angiotensin-converting enzyme (ACE) inhibitor treatment increases the lifespan of long-lived F1 male mice. Age (Dordr) 38, 379–391, doi: 10.1007/s11357-016-9948-4 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Evason K, Huang C, Yamben I, Covey DF & Kornfeld K Anticonvulsant medications extend worm life-span. Science (New York, N.Y 307, 258–262, doi: 10.1126/science.1105299 (2005). [DOI] [PubMed] [Google Scholar]
- 93.Evason K, Collins JJ, Huang C, Hughes S & Kornfeld K Valproic acid extends Caenorhabditis elegans lifespan. Aging cell 7, 305–317, doi: 10.1111/j.1474-9726.2008.00375.x (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Admasu TD et al. Drug Synergy Slows Aging and Improves Healthspan through IGF and SREBP Lipid Signaling. Developmental cell 47, 67–79 e65, doi: 10.1016/j.devcel.2018.09.001 (2018). [DOI] [PubMed] [Google Scholar]
- 95.Castillo-Quan JI et al. Lithium Promotes Longevity through GSK3/NRF2-Dependent Hormesis. Cell reports 15, 638–650, doi: 10.1016/j.celrep.2016.03.041 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Slack C et al. The Ras-Erk-ETS-Signaling Pathway Is a Drug Target for Longevity. Cell 162, 72–83, doi: 10.1016/j.cell.2015.06.023 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Castillo-Quan JI et al. A triple drug combination targeting components of the nutrient-sensing network maximizes longevity. Proceedings of the National Academy of Sciences of the United States of America 116, 20817–20819, doi: 10.1073/pnas.1913212116 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Danilov A et al. Selective anticancer agents suppress aging in Drosophila. Oncotarget 4, 1507–1526, doi: 10.18632/oncotarget.1272 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Bustos V & Partridge L Good Ol' Fat: Links between Lipid Signaling and Longevity. Trends Biochem Sci 42, 812–823, doi: 10.1016/j.tibs.2017.07.001 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Johnson AA & Stolzing A The role of lipid metabolism in aging, lifespan regulation, and age-related disease. Aging cell 18, e13048, doi: 10.1111/acel.13048 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Hou NS & Taubert S Function and Regulation of Lipid Biology in Caenorhabditis elegans Aging. Front Physiol 3, 143, doi: 10.3389/fphys.2012.00143 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Huang X et al. Reducing signs of aging and increasing lifespan by drug synergy. Aging cell 12, 652–660, doi: 10.1111/acel.12090 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Huang X, Leggas M & Dickson RC Drug synergy drives conserved pathways to increase fission yeast lifespan. PLoS One 10, e0121877, doi: 10.1371/journal.pone.0121877 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Hubbard BP & Sinclair DA Small molecule SIRT1 activators for the treatment of aging and age-related diseases. Trends Pharmacol Sci 35, 146–154, doi: 10.1016/j.tips.2013.12.004 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Mercken EM et al. SRT2104 extends survival of male mice on a standard diet and preserves bone and muscle mass. Aging cell 13, 787–796, doi: 10.1111/acel.12220 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Mitchell SJ et al. The SIRT1 activator SRT1720 extends lifespan and improves health of mice fed a standard diet. Cell reports 6, 836–843, doi: 10.1016/j.celrep.2014.01.031 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Minor RK et al. SRT1720 improves survival and healthspan of obese mice. Sci Rep 1, 70, doi: 10.1038/srep00070 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Palliyaguru DL et al. Combining a High Dose of Metformin With the SIRT1 Activator, SRT1720, Reduces Life Span in Aged Mice Fed a High-Fat Diet. The journals of gerontology. Series A, Biological sciences and medical sciences 75, 2037–2041, doi: 10.1093/gerona/glaa148 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Shen Z, Hinson A, Miller RA & Garcia GG Cap-independent translation: A shared mechanism for lifespan extension by rapamycin, acarbose, and 17alpha-estradiol. Aging cell 20, e13345, doi: 10.1111/acel.13345 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Tiku V et al. Small nucleoli are a cellular hallmark of longevity. Nature communications 8, 16083, doi: 10.1038/ncomms16083 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Annibal A et al. Regulation of the one carbon folate cycle as a shared metabolic signature of longevity. Nature communications 12, 3486, doi: 10.1038/s41467-021-23856-9 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Qiao HH et al. An efficient and multiple target transgenic RNAi technique with low toxicity in Drosophila. Nature communications 9, 4160, doi: 10.1038/s41467-018-06537-y (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Port F & Bullock SL Augmenting CRISPR applications in Drosophila with tRNA-flanked sgRNAs. Nature methods 13, 852–854, doi: 10.1038/nmeth.3972 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Parkhitko AA et al. Cross-species identification of PIP5K1-, splicing- and ubiquitin-related pathways as potential targets for RB1-deficient cells. PLoS Genet 17, e1009354, doi: 10.1371/journal.pgen.1009354 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Norris AD, Gracida X & Calarco JA CRISPR-mediated genetic interaction profiling identifies RNA binding proteins controlling metazoan fitness. eLife 6, doi: 10.7554/eLife.28129 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Horvath S DNA methylation age of human tissues and cell types. Genome Biol 14, R115, doi: 10.1186/gb-2013-14-10-r115 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Fahy GM et al. Reversal of epigenetic aging and immunosenescent trends in humans. Aging cell 18, e13028, doi: 10.1111/acel.13028 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Kaeberlein TL et al. Lifespan extension in Caenorhabditis elegans by complete removal of food. Aging cell 5, 487–494, doi: 10.1111/j.1474-9726.2006.00238.x (2006). [DOI] [PubMed] [Google Scholar]
- 119.Hoffmann M et al. MICS-1 interacts with mitochondrial ATAD-3 and modulates lifespan in C. elegans. Experimental gerontology 47, 270–275, doi: 10.1016/j.exger.2011.12.011 (2012). [DOI] [PubMed] [Google Scholar]
- 120.Zimmerman SM, Hinkson IV, Elias JE & Kim SK Reproductive Aging Drives Protein Accumulation in the Uterus and Limits Lifespan in C. elegans. PLoS Genet 11, e1005725, doi: 10.1371/journal.pgen.1005725 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Jia K, Albert PS & Riddle DL DAF-9, a cytochrome P450 regulating C. elegans larval development and adult longevity. Development 129, 221–231, doi: 10.1242/dev.129.1.221 (2002). [DOI] [PubMed] [Google Scholar]
- 122.Snell TW, Johnston RK, Rabeneck B, Zipperer C & Teat S Joint inhibition of TOR and JNK pathways interacts to extend the lifespan of Brachionus manjavacas (Rotifera). Experimental gerontology 52, 55–69, doi: 10.1016/j.exger.2014.01.022 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Roux AE, Quissac A, Chartrand P, Ferbeyre G & Rokeach LA Regulation of chronological aging in Schizosaccharomyces pombe by the protein kinases Pka1 and Sck2. Aging cell 5, 345–357, doi: 10.1111/j.1474-9726.2006.00225.x (2006). [DOI] [PubMed] [Google Scholar]
- 124.Shaposhnikov MV et al. Molecular mechanisms of exceptional lifespan increase of Drosophila melanogaster with different genotypes after combinations of pro-longevity interventions. Commun Biol 5, 566, doi: 10.1038/s42003-022-03524-4 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Abergel R, Livshits L, Shaked M, Chatterjee AK & Gross E Synergism between soluble guanylate cyclase signaling and neuropeptides extends lifespan in the nematode Caenorhabditis elegans. Aging cell 16, 401–413, doi: 10.1111/acel.12569 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Prasanth MI, Gayathri S, Bhaskar JP, Krishnan V & Balamurugan K Analyzing the Synergistic Effects of Antioxidants in Combating Photoaging Using Model Nematode, Caenorhabditis elegans. Photochem Photobiol 96, 139–147, doi: 10.1111/php.13167 (2020). [DOI] [PubMed] [Google Scholar]
- 127.Imanikia S, Hylands P & Sturzenbaum SR The double mutation of cytochrome P450's and fatty acid desaturases affect lipid regulation and longevity in C. elegans. Biochem Biophys Rep 2, 172–178, doi: 10.1016/j.bbrep.2015.06.007 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Westfall S, Lomis N & Prakash S Longevity extension in Drosophila through gut-brain communication. Sci Rep 8, 8362, doi: 10.1038/s41598-018-25382-z (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Maures TJ, Greer EL, Hauswirth AG & Brunet A The H3K27 demethylase UTX-1 regulates C. elegans lifespan in a germline-independent, insulin-dependent manner. Aging cell 10, 980–990, doi: 10.1111/j.1474-9726.2011.00738.x (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Frankowski H et al. Dimethyl sulfoxide and dimethyl formamide increase lifespan of C. elegans in liquid. Mechanisms of ageing and development 134, 69–78, doi: 10.1016/j.mad.2012.10.002 (2013). [DOI] [PubMed] [Google Scholar]
- 131.Ng LT et al. Lifespan and healthspan benefits of exogenous H2S in C. elegans are independent from effects downstream of eat-2 mutation. NPJ aging and mechanisms of disease 6, 6, doi: 10.1038/s41514-020-0044-8 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Gusarov I et al. Bacterial nitric oxide extends the lifespan of C. elegans. Cell 152, 818–830, doi: 10.1016/j.cell.2012.12.043 (2013). [DOI] [PubMed] [Google Scholar]
- 133.Kim SH, Kim BK, Park S & Park SK Phosphatidylcholine Extends Lifespan via DAF-16 and Reduces Amyloid-Beta-Induced Toxicity in Caenorhabditis elegans. Oxid Med Cell Longev 2019, 2860642, doi: 10.1155/2019/2860642 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Lakowski B & Hekimi S Determination of life-span in Caenorhabditis elegans by four clock genes. Science (New York, N.Y 272, 1010–1013, doi: 10.1126/science.272.5264.1010 (1996). [DOI] [PubMed] [Google Scholar]
- 135.Chandler-Brown D et al. Sorbitol treatment extends lifespan and induces the osmotic stress response in Caenorhabditis elegans. Frontiers in genetics 6, 316, doi: 10.3389/fgene.2015.00316 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Siebold AP et al. Polycomb Repressive Complex 2 and Trithorax modulate Drosophila longevity and stress resistance. Proceedings of the National Academy of Sciences of the United States of America 107, 169–174, doi: 10.1073/pnas.0907739107 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Meng F, Li J, Rao Y, Wang W & Fu Y Gengnianchun Extends the Lifespan of Caenorhabditis elegans via the Insulin/IGF-1 Signalling Pathway. Oxid Med Cell Longev 2018, 4740739, doi: 10.1155/2018/4740739 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Yang ZZ et al. Lonicera japonica extends lifespan and healthspan in Caenorhabditis elegans. Free Radic Biol Med 129, 310–322, doi: 10.1016/j.freeradbiomed.2018.09.035 (2018). [DOI] [PubMed] [Google Scholar]
- 139.Wilhelm T et al. Neuronal inhibition of the autophagy nucleation complex extends life span in post-reproductive C. elegans. Genes & development 31, 1561–1572, doi: 10.1101/gad.301648.117 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.