Skip to main content

OPINION article

Front. Aging, 04 October 2022
Sec. Genetics, Genomics and Epigenomics of Aging
Volume 3 - 2022 | https://doi.org/10.3389/fragi.2022.1016962

Fly immunity comes of age: The utility of Drosophila as a model for studying variation in immunosenescence

  • Institute of Immunology and Infection Research, University of Edinburgh, Edinburgh, United Kingdom

1 Introduction

1.1 The challenge of understanding individual variation in immunosenescence

The global demographic is shifting towards an aged population (United Nations, 2019). Yet, it is our oldest members of society that are the most vulnerable to infectious disease, reflecting immune system ageing, or “immunosenescence” (Glynn and Moss, 2020; Chen et al., 2021). Immunosenescence has two defining features: heightened susceptibility to pathogens, and increased systemic, basal inflammation (Franceschi, 2000; Zerofsky et al., 2005; Benayoun et al., 2019), which is strongly linked to the development of age-related diseases such as neurodegeneration and cancer (Franceschi et al., 2018; Rea, 2018). Understanding the mechanisms of immunosenescence could offer therapeutic potential in reducing age-associated morbidity (Neves and Sousa-Victor, 2020; Borgoni et al., 2021). However, individuals do not age at the same rate, with variation arising from genotype (Burger and Promislow, 2006; Belsky et al., 2015), including sex (Sampathkumar et al., 2020; Xirocostas et al., 2020; Bronikowski et al., 2022), and from the environment. While most mechanistic biology research ignores this diversity, largely due to the practicality of using single genotypes and sexes, individual variation will determine the response to therapeutics targeting ageing pathologies (Meyer et al., 2013; Fuselli, 2019). Here, we champion the use of Drosophila melanogaster as an ideal model for the early stages of an approach that aims to leverage population variation to understand immunosenescence and how to treat it. We briefly describe the rich history of this model system in the fields of the biology of ageing, biogerontology and immunology, and their conceptual and technical convergence on the fly model. We outline the unique potential of Drosophila in elucidating variation in immunosenescence, its underlying mechanisms and its treatments.

1.2 Drosophila’s central role in formulating evolutionary theories of ageing—why do organisms age?

Drosophila have proven pivotal in the synthesis and empirical testing of the evolutionary theories of ageing, which provide a conceptual framework within which to understand the biology of ageing. Experimental evolution experiments highlighted a trade-off between longevity and early-life fecundity (Rose and Charlesworth, 1981; Luckinbill and Clare, 1985; Rose, 1991). Such fitness trade-offs are commonly seen in insects (Flatt et al., 2013) and give credence to Williams’ theory of antagonistic pleiotropy (Williams, 1957). These trade-offs may rely on resource allocation between germline and “disposable” soma (Kirkwood, 1977), and are not a universal, but rather a contextual, quid pro quo that depends on environmental factors. The range of phenotypic expression of a genotype across varying environments is called phenotypic plasticity. It is an important source of natural variation for life history traits. This has been demonstrated in multiple settings, such as temperature and commensal status, using Drosophila (e.g., Lee et al., 2019; Huang et al., 2020). An influential evolutionary theory, the resource reallocation hypothesis, suggests that natural selection may favour phenotypically plastic responses in allocation, prioritising somatic maintenance during scarcity, and reproduction in replete settings (Shanley and Kirkwood, 2000; Regan, 2020). Drosophila has offered an easy-to-manipulate empirical platform to understand these phenotypically plastic responses to nutritional resources, and their associated trade-offs (Lee et al., 2008; Jensen et al., 2015; Lee, 2015; Zanco et al., 2021). These studies have served to shape our understanding of key evolutionary concepts on the consequence of reproductive investment, and phenotypically plastic responses to the environment, in determining rates of ageing.

1.3 Drosophila’s central role in biogerontology—how can we slow down ageing?

The goal of Biogerontology, born out of evolutionary biology, is to enhance healthy ageing by limiting age-associated multi-morbidity through therapeutic intervention. Key to this goal is a thorough understanding of the underlying molecular mechanisms that govern the ageing process. Ageing has been studied using flies for over a century (Hyde, 1913; Piper and Partridge, 2018). One of the observations fundamental to modern-day biogerontology was the remarkably beneficial effects of dietary restriction on ageing across taxa. The use of invertebrate model organisms, such as flies and worms, has given insight into the dietary macronutrients, genes and pathways underpinning this effect (Grandison et al., 2009; Piper and Partridge, 2018). These studies have not only implicated environment-sensing pathways like IIS (Clancy et al., 2001; Tatar et al., 2001), and TOR (Vellai et al., 2003; Kapahi, 2004), but also numerous longevity-associated pathways including Toll, Ras-ERK-ETS, AMPK, and Myc pathways (Piper and Partridge, 2018). The extension of lifespan through targeted manipulation of pathways such as TOR was initially demonstrated in invertebrates (Vellai et al., 2003), including Drosophila (Bjedov et al., 2010), before it was shown to be effective in mammals (Harrison et al., 2009).

1.4 Drosophila’s central role in immunology—the genetics of innate immunity

The famous intersection of genetic tractability with immunology following the identification of Toll as an immune signalling molecule in Drosophila (Lemaitre et al., 1996), and later characterisation of mammalian Toll-like pathogen-recognition receptors (Medzhitov et al., 1997), sealed Drosophila’s fate as an indispensable model system in innate immunity research. Indeed, there are multiple conserved immune components of the model. Immune signalling culminates in expression of antimicrobial peptides (AMPs) and immune effectors via conserved NF-kB-like transcription factors (Lemaitre and Hoffmann, 2007; Buchon et al., 2014). Haematopoiesis, controlled by conserved transcription factors (Evans et al., 2003), gives rise to macrophage-like hemocytes which share a common superfamily of phagocytic receptors with humans (Kurucz et al., 2007). Epithelial immunity, such as the gut, is governed by conserved immune signalling, regulates commensal interactions, and is structurally similar to that of humans (Liu et al., 2017).

1.5 Power to the fly: Drosophila as a nexus for immunity, evolution and biogerontology to understand variation in immunosenescence

The rich history of Drosophila as a model system, spanning more than a century, has produced biological insights too numerous to mention. Here, we have briefly detailed the contributions of the model to three, interconnected disciplines. The power, we argue, in using Drosophila as a model in immunosenescence, which shares many similarities with that of mammals and has been reviewed elsewhere (Flatt and Garschall, 2018; Min and Tatar, 2018; Sciambra and Chtarbanova, 2021), comes from the conceptual and technical convergence of these fields on the system (Figure 1). Coupling the genetic and technical approaches possible in Drosophila with an evolutionary biology approach to address different forms of variation in immune responses will not only provide mechanistic insights into immunosenescence, but will also guide our ultimate goal of treating it.

FIGURE 1
www.frontiersin.org

FIGURE 1. Drosophila’s rich history in the fields of evolutionary biology of ageing, biogerontology and immunity have endowed the model with multiple conceptual approaches and techniques. The functional genetic approaches, ease of environmental and sex manipulation render it invaluable in understanding natural variation in immunosenescence and its potential treatments. Created with BioRender.com.

1.5.1 Genetic variation

Genetic and phenotypic heterogeneity in immune responses are ubiquitous across species. Panels of fully-sequenced, isogenic lines derived from wild-caught Drosophila (Mackay, 2012; Grenier et al., 2015), or from extensive recombination (King, 2012), are an important community resource for understanding the genetic bases of complex phenotypic traits. These lines have been used to explore natural variation in, and genetics underpinning, susceptibility to pathogens. Susceptibility is determined by both resistance, which describes the ability to control pathogen burden (Wang et al., 2017; Palmer, 2018; Chapman et al., 2020) and disease tolerance (Howick and Lazzaro, 2017), which describes the ability to cope with infection and represents a less well-understood immune strategy (Soares et al., 2017; Schneider, 2021). Drosophila offers an ideal model to address the elusive processes determining disease tolerance, as they offer readily measurable fitness metrics and genetic tractability (e.g., Kutzer and Armitage, 2016; Duneau et al., 2017a; Gupta and Vale, 2017). Assessing variation in disease tolerance to Providencia rettgeri has implicated several biological processes, such as genes related to the endoplasmic reticulum (Howick and Lazzaro, 2017), which was further linked to tolerance in a transcriptomic analysis of infection response (Troha et al., 2018). These studies have granted some initial insight into the physiology underpinning this relatively elusive immune strategy, and genetic variation therein.

Use of isogenic panels could be extended to explore natural variation and plasticity in immunosenescence and the immune strategies affected by age. The role for declines in disease tolerance, and its variation, in driving immunosenescence is currently a black box, largely due to our lack of understanding of the mechanisms underpinning tolerance. There is a dearth of genotypically high-powered experiments assessing genetic variation in age-related immune dysregulation, despite prevalent genotype-by-age interactions (Lesser et al., 2006; Felix et al., 2012).

By performing genome wide association analysis (GWAS; e.g., King, 2012; Mackay, 2012) when assessing such variation in immunosenescence among the isogenic panels, candidate genes and pathways could potentially be identified, and ultimately validated. The ability to conditionally manipulate virtually any gene in the fly genome, through the spatiotemporal control of transgene expression via binary genetic systems (Brand and Perrimon, 1993; Osterwalder et al., 2001), makes validation of candidate genes possible in high throughput. These tools grant the model relative ease in assessing canonical and non-canonical components of immune defence and their impact on immunosenescence over ageing. Additionally, Drosophila are tractable to reverse genetic knock-out approaches using CRISPR/Cas9 technology, as has been demonstrated in the functional assessment of individual AMPs (Hanson et al., 2019). Recently, the potential of such CRISPR lines in the functional assessment of AMPs over ageing has been demonstrated. While ageing is associated with expansion of the AMP repertoire and loss of specificity, possibly indicative of immune dysregulation (Shit et al., 2022), their collective role in regulating microbiome dysbiosis was shown to be a critical determinant of lifespan (Hanson and Lemaitre, 2022), the control of which has previously been linked to lifespan and ageing pathologies (Guo et al., 2014; Clark et al., 2015; Li et al., 2016) Finally, fluorescent labelling of immune tissues, such as hemocytes (e.g., Sanchez Bosch et al., 2019; Krejčová, 2019; Chakrabarti and Visweswariah, 2020; Kierdorf, 2020; Coates, 2021), opens the possibility to assess their contribution to immune, and systemic, ageing.

Flies are highly amenable to experimental evolution, which when coupled with genomic sequencing, allows inferences to be made about the genetic response to a dominant selection pressure, such as delayed reproduction or longevity (McHugh and Burke, 2022). This technique can be used to explore trade-offs between immunity and life-history traits (McKean et al., 2008; Shahrestani et al., 2021), or lack thereof (Faria, 2015; Gupta, 2016). In lines selected for longevity, an improvement of measurable immune responses was observed (Fabian, 2018), whereas selection for resistance against a specific fungal pathogen was costly to lifespan (Shahrestani et al., 2021). It is possible, and remains to be tested, that improved early immunity has detrimental effects on the rate of immunosenescence, ultimately affecting ageing.

1.5.2 Plasticity in immunity

Immunology research has historically focused on the genetics underpinning immune responses within a controlled environment, but rarely assesses reaction norms under varying conditions (Martin, 2021). Drosophila has offered insight into such immune plasticity, including genotype-by-temperature (Lazzaro et al., 2008), and genotype-by-diet interactions using isogenic lines. The observed effects of diet on resistance (Unckless et al., 2015) underscore the importance of nutrition in determining immune responses (for e.g., Galenza, 2016; Ponton et al., 2020). Similarly, disease tolerance plasticity was observed in a diet- and pathogen-specific manner (Howick and Lazzaro, 2014; Kutzer and Armitage, 2016). Through the assessment of genotype-by-environment interactions, these investigations have highlighted an intrinsic link between immunity and metabolism, as has been confirmed elsewhere, for example, the induction of immune-response genes downstream of IIS and TOR nutrient-sensing pathways independently of Toll or Imd signalling (Becker et al., 2010; Varma et al., 2014). Thus, environmental factors are an important source of variation in immunity, as predicted by evolutionary theory.

Little is known about phenotypic plasticity in the manifestation of immunosenescence. Yet, the question “what kind of environments accelerate or slow down immunosenescence?” is crucial to an evolutionary biology approach in treating immune decline. Plasticity in immune responses in the face of varying environments early in life (Lazzaro and Little, 2009; Leech, 2019) is likely to have consequences on immunosenescence. While such plasticity is often ignored, or difficult to address, as in mouse studies (Martin, 2021), assessing the expression of immune responses across differing environments in flies is relatively tenable (McKean and Nunney, 2005; Lazzaro et al., 2008; Unckless et al., 2015).

Pharmacological intervention represents a highly relevant example of environmental manipulation, and we have barely scratched the surface in understanding genetic variation underpinning responses to geroprotective treatments (but see Lind, 2017; Rohde et al., 2021). Through robust genetic validation, flies have provided mechanistic insight into the action of multiple potential geroprotective drugs (for e.g., Bjedov et al., 2010; Slack et al., 2015; Castillo-Quan et al., 2019). Drosophila is the ideal candidate model to move towards high-powered studies addressing variation in response to either environmental (e.g., dietary restriction) or pharmacological (e.g., therapeutic mTOR attenuation) interventions targeting ageing, including immunosenescence. The crucial next step in biogerontology is to capture individual variation in responses to therapeutics, which, in contrast to mammalian systems, can be tested with relative ease in Drosophila.

1.5.3 Sex bias in immunity

Sex differences in immune responses, which have been observed across taxa (Kelly et al., 2018; Metcalf et al., 2020), are prevalent in Drosophila infectious disease models (Belmonte et al., 2020). Sex differences in the immune response appear contextual in Drosophila (Belmonte et al., 2020), mirroring the lack of ubiquitous sex differences across taxa (Kelly et al., 2018). Sex differences in infection outcome are likely to be dictated by a complex interplay of environmental factors, pathogen-specific consequences of infection, and the different life-histories of the sexes. Notwithstanding this complexity, fruit flies have offered mechanistic insight into pathogen-specific, dimorphic responses to infection, such as sex differences in the Toll pathway (D. Duneau et al., 2017a; Shahrestani et al., 2018), and sex-by-genotype interactions determining viral load and transmission potential of Drosophila C virus (Siva-Jothy and Vale, 2021).

The inclusion of both sexes when using isogenic, transgenic or outbred Drosophila lines will invariably provide insight into sex differences in immunosenescence. In particular, examining dimorphism across genetic panels could address the interaction of sex and ageing on immunosenescence. Furthermore, the ability to remove the gonads prior to next generation sequencing facilitates the examination of somatic, sex differential gene expression, as was done to compare the immune response to P. rettgeri (Duneau et al., 2017b). Coupling the cell autonomous sex determination system of Drosophila (Salz and Erickson, 2010) with their robust genetic toolkit allows the manipulation of sex in a tissue-specific manner, as has been demonstrated in exploration of sexual dimorphisms in physiology (Hudry et al., 2016; Millington et al., 2021) and in the infection response and pathology of the ageing gut (Regan et al., 2016). This offers an unparalleled system to tease apart how tissue-specific and systemic sex differences contribute to individual variation in immunosenescence. Considering the conserved dimorphism in both immune function and responses to drugs that delay ageing (Bjedov et al., 2010; Arriola Apelo and Lamming, 2016; Bitto et al., 2016; Regan et al., 2016; Regan et al., 2021; Strong et al., 2020); Drosophila genetic techniques could be leveraged for the early stages of developing sex-optimised drug treatments for immunosenescence.

2 Conclusion

Immune decline over ageing is linked to age-associated morbidity through increased pathogen susceptibility and dysregulated inflammation. Yet, immunosenescence remains relatively elusive where individual-to-individual variation in the onset and extent is virtually unknown, potentially determining responses to treatments targeting the decline. Steps to capturing population diversity in immune responses, arising from genotype, phenotypic plasticity and sex, discussed here, would provide mechanistic insight if applied to immunosenescence. We argue that Drosophila is uniquely situated to address the challenge of variation in ageing. Its vast history as a model in the fields of evolutionary biology, biogerontology, and immunity offers a foundation of knowledge about the species and a matchless system of tractable genetics, statistical power, and environmental manipulation. Using Drosophila to initiate an evolutionary biology approach to immunosenescence would address population diversity and, ultimately, aid the development of personalised therapeutics.

Author contributions

M-KC and JR conceived the article. M-KC wrote the manuscript with contribution and editing from JR. Both authors contributed to the article and approved the submitted version.

Funding

M-KC and JR were funded by The University of Edinburgh.

Acknowledgments

Thanks to David F. Duneau for valuable comments on the manuscript, and to members of the Regan lab for discussion.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Arriola Apelo, S. I., and Lamming, D. W. (2016). Rapamycin: An InhibiTOR of aging emerges from the soil of easter island. J. Gerontol. A Biol. Sci. Med. Sci. 71 (7), 841–849. doi:10.1093/gerona/glw090

PubMed Abstract | CrossRef Full Text | Google Scholar

Becker, T., Loch, G., Beyer, M., Zinke, I., Aschenbrenner, A. C., Carrera, P., et al. (2010). FOXO-dependent regulation of innate immune homeostasis. Nature 463 (7279), 369–373. doi:10.1038/nature08698

PubMed Abstract | CrossRef Full Text | Google Scholar

Belmonte, R. L., Corbally, M. K., Duneau, D. F., and Regan, J. C. (2020). Sexual dimorphisms in innate immunity and responses to infection in Drosophila melanogaster. Front. Immunol. 10, 3075. doi:10.3389/FIMMU.2019.03075/BIBTEX

PubMed Abstract | CrossRef Full Text | Google Scholar

Belsky, D. W., Caspi, A., Houts, R., Cohen, H. J., Corcoran, D. L., Danese, A., et al. (2015). Quantification of biological aging in young adults. Proc. Natl. Acad. Sci. U. S. A. 112 (30), E4104–E4110. doi:10.1073/pnas.1506264112

PubMed Abstract | CrossRef Full Text | Google Scholar

Benayoun, B. A., Pollina, E. A., Priya Singh, P., Mahmoudi, S., Harel, I., Casey, K. M., et al. (2019). Remodeling of epigenome and transcriptome landscapes with aging in mice reveals widespread induction of inflammatory responses. Genome Res. 29, 697–709. doi:10.1101/gr.240093.118

PubMed Abstract | CrossRef Full Text | Google Scholar

Bitto, A., Ito, T. K., Pineda, V. V., LeTexier, N. J., Huang, H. Z., Sutlief, E., et al. (2016). Transient rapamycin treatment can increase lifespan and healthspan in middle-aged mice. eLife 5, e16351. doi:10.7554/eLife.16351

PubMed Abstract | CrossRef Full Text | Google Scholar

Bjedov, I., Toivonen, J. M., Kerr, F., Slack, C., Jacobson, J., Foley, A., et al. (2010). Mechanisms of life span extension by rapamycin in the fruit fly Drosophila melanogaster, ” Cell metabolism. Cell Metab. 11 (1), 35–46. doi:10.1016/J.CMET.2009.11.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Borgoni, S., Kudryashova, K. S., Burka, K., and de Magalhaes, J. P. (2021). Targeting immune dysfunction in aging. Ageing Res. Rev. 70, 101410. doi:10.1016/j.arr.2021.101410

PubMed Abstract | CrossRef Full Text | Google Scholar

Brand, A. H., and Perrimon, N. (1993). Targeted gene expression as a means of altering cell fates and generating dominant phenotypes. Development 118 (2), 401–415. doi:10.1242/dev.118.2.401

PubMed Abstract | CrossRef Full Text | Google Scholar

Bronikowski, A. M., Meisel, R. P., Peggy, R., Biga, R., Walters, J. R., Mank, J. E., et al. (2022). Sex-specific aging in animals: Perspective and future directions. Aging Cell 21, e13542e13542. doi:10.1111/acel.13542

PubMed Abstract | CrossRef Full Text | Google Scholar

Buchon, N., Silverman, N., and Cherry, S. (2014). Immunity in Drosophila melanogaster — From microbial recognition to whole-organism physiology. Nat. Rev. Immunol. 14, 796–810. doi:10.1038/nri3763

PubMed Abstract | CrossRef Full Text | Google Scholar

Burger, J. M. S., and Promislow, D. E. L. (2006). Are functional and demographic senescence genetically independent? Exp. Gerontol. 41 (11), 1108–1116. doi:10.1016/j.exger.2006.08.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Castillo-Quan, J. I., Tain, L. S., Kinghorn, K. J., Li, L., Gronke, S., Hinze, Y., et al. (2019). A triple drug combination targeting components of the nutrient-sensing network maximizes longevity. Proc. Natl. Acad. Sci. U. S. A. 116 (42), 20817–20819. doi:10.1073/pnas.1913212116

PubMed Abstract | CrossRef Full Text | Google Scholar

Chakrabarti, S., and Visweswariah, S. S. (2020). Intramacrophage ROS primes the innate immune system via JAK/STAT and Toll activation. Cell Rep. 33 (6), 108368. doi:10.1016/J.CELREP.2020.108368

PubMed Abstract | CrossRef Full Text | Google Scholar

Chapman, J. R., Dowell, M. A., Chan, R., and Unckless, R. L. (2020). The genetic basis of natural variation in Drosophila melanogaster immune defense against Enterococcus faecalis. Genes 11, 234. doi:10.3390/GENES11020234

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, Y., Klein, S. L., Garibaldi, B. T., Li, H., Wu, C., Osevala, N. M., et al. (2021). Aging in COVID-19: Vulnerability, immunity and intervention. Ageing Res. Rev. 65, 101205. doi:10.1016/j.arr.2020.101205

PubMed Abstract | CrossRef Full Text | Google Scholar

Clancy, D. J., Gems, D., Harshman, L. G., Oldham, S., Stocker, H., Hafen, E., et al. (2001). Extension of life-span by loss of CHICO, a Drosophila insulin receptor substrate protein,” Science. Am. Assoc. Adv. Sci., 292, 104–106. doi:10.1126/SCIENCE.1057991/ASSET/04E1720A-7F71-4F12-A9DB-00508D22FF36/ASSETS/GRAPHIC/SE1219311004

PubMed Abstract | CrossRef Full Text | Google Scholar

Clark, R. I., Salazar, A., Yamada, R., Fitz-Gibbon, R., Morselli, R., Alcaraz, J., et al. (2015). Distinct shifts in microbiota composition during Drosophila aging impair intestinal function and drive mortality. Cell Rep. 12, 1656–1667. doi:10.1016/J.CELREP.2015.08.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Coates, J. A. (2021). Identification of functionally distinct macrophage subpopulations in drosophila eLife 10. doi:10.7554/ELIFE.58686

CrossRef Full Text | Google Scholar

Duneau, D., Ferdy, J. B., Revah, J., Kondolf, H., Ortiz, G. A., Lazzaro, P. B., et al. (2017a). Stochastic variation in the initial phase of bacterial infection predicts the probability of survival in D. melanogaster. eLife, 6, 28298. doi:10.7554/ELIFE.28298

PubMed Abstract | CrossRef Full Text | Google Scholar

Duneau, D., Kondolf, H. C., Im, J. H., Ortiz, G. A., Chow, C., Fox, M. A., et al. (2017b). The Toll pathway underlies host sexual dimorphism in resistance to both Gram-negative and Gram-positive bacteria in mated Drosophila. BMC Biol. 15 (1), 124. doi:10.1186/s12915-017-0466-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Evans, C. J., Hartenstein, V., and Banerjee, U. (2003). Thicker than blood: Conserved mechanisms in Drosophila and vertebrate hematopoiesis. Dev. Cell 5 (5), 673–690. doi:10.1016/S1534-5807(03)00335-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Fabian, D. K. (2018). Evolution of longevity improves immunity in Drosophila. Evol. Lett. 2, 567–579. doi:10.1002/evl3.89

PubMed Abstract | CrossRef Full Text | Google Scholar

Faria, V. G. (2015). Evolution of Drosophila resistance against different pathogens and infection routes entails no detectable maintenance costs. Evol. Int. J. Org. Evol 69, 2799–2809. doi:10.1111/EVO.12782

PubMed Abstract | CrossRef Full Text | Google Scholar

Felix, T. M., Hughes, K. A., Stone, E. A., Drnevich, J. M., and Leips, J. (2012). Age-specific variation in immune response in Drosophila melanogaster has a genetic basis. Genetics 191 (3), 989–1002. doi:10.1534/genetics.112.140640

PubMed Abstract | CrossRef Full Text | Google Scholar

Flatt, T., Stone, E. A., Drnevich, J. M., and Leips, J. (2013). Life-history evolution and the polyphenic regulation of somatic maintenance and survival. Q. Rev. Biol. 88, 185–218. doi:10.1086/671484/ASSET/IMAGES/LARGE/FG6_ONLINE

PubMed Abstract | CrossRef Full Text | Google Scholar

Flatt, T., and Garschall, K. (2018). The interplay between immunity and aging in Drosophila. F1000Research 7, 1. doi:10.12688/F1000RESEARCH.13117.1

PubMed Abstract | CrossRef Full Text | Google Scholar

Franceschi, C. (2000). Inflamm-aging. An evolutionary perspective on immunosenescence,” Annals of the New York Academy of Sciences. Ann N Y Acad Sci 908, 244–254. doi:10.1111/J.1749-6632.2000.TB06651.X

PubMed Abstract | CrossRef Full Text | Google Scholar

Franceschi, C., Garagnani, P., Parini, P., Giuliani, C., and Santoro, A. (2018). Inflammaging: A new immune–metabolic viewpoint for age-related diseases. Nat. Rev. Endocrinol. 14 (10), 576–590. doi:10.1038/s41574-018-0059-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Fuselli, S. (2019). Beyond drugs: The evolution of genes involved in human response to medications. Proc. Biol. Sci. 286, 20191716. doi:10.1098/rspb.2019.1716

PubMed Abstract | CrossRef Full Text | Google Scholar

Galenza, A (2016). Glucose modulates Drosophila longevity and immunity independent of the microbiota. Biol. open 5, 165–173. doi:10.1242/bio.015016

PubMed Abstract | CrossRef Full Text | Google Scholar

Glynn, J. R., and Moss, P. A. H. (2020). Systematic analysis of infectious disease outcomes by age shows lowest severity in school-age children. Sci. Data 7 (1), 329. doi:10.1038/s41597-020-00668-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Grandison, R. C., Piper, M. D. W., and Partridge, L. (2009). Amino acid imbalance explains extension of lifespan by dietary restriction in Drosophila. Nature 462 (7276), 1061–1064. doi:10.1038/NATURE08619

PubMed Abstract | CrossRef Full Text | Google Scholar

Grenier, J. K., Arguello, J. R., Moreira, M. C., Gottipati, S., Mohammed, J., Hackett, S. R., et al. (2015). Global diversity lines - a five-continent reference panel of sequenced Drosophila melanogaster strains. G3 5 (4), 593–603. doi:10.1534/g3.114.015883

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, L., Karpac, J., Tran, S. L., and Jasper, H. (2014). PGRP-SC2 promotes gut immune homeostasis to limit commensal dysbiosis and extend lifespan. Cell 156 (0), 109–122. doi:10.1016/J.CELL.2013.12.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Gupta, V. (2016). No apparent cost of evolved immune response in Drosophila melanogaster. Evol. Int. J. Org. Evol, 70, 934–943. doi:10.1111/EVO.12896

PubMed Abstract | CrossRef Full Text | Google Scholar

Gupta, V., and Vale, P. F. (2017). Nonlinear disease tolerance curves reveal distinct components of host responses to viral infection. R. Soc. Open Sci. 4, 170342. doi:10.1098/RSOS.170342

PubMed Abstract | CrossRef Full Text | Google Scholar

Hanson, M. A., Dostalova, A., Ceroni, C., Poidevin, M., Kondo, S., and Lemaitre, B. (2019). Synergy and remarkable specificity of antimicrobial peptides in vivo using a systematic knockout approach. eLife 8, e44341. doi:10.7554/eLife.44341

PubMed Abstract | CrossRef Full Text | Google Scholar

Hanson, M. A., and Lemaitre, B. (2022).Antimicrobial peptides do not directly contribute to aging in Drosophila, but improve lifespan by preventing dysbiosis. bioRxiv. doi:10.1101/2022.08.24.504952

CrossRef Full Text | Google Scholar

Harrison, D. E., Strong, R., Sharp, Z. D., Nelson, J. F., Astle, C. M., Flurkey, K., et al. (2009). Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature 460 (7253), 392–395. doi:10.1038/nature08221

PubMed Abstract | CrossRef Full Text | Google Scholar

Howick, V. M., and Lazzaro, B. P. (2014). Genotype and diet shape resistance and tolerance across distinct phases of bacterial infection. BMC Evol. Biol. 14 (1), 56. doi:10.1186/1471-2148-14-56

PubMed Abstract | CrossRef Full Text | Google Scholar

Howick, V. M., and Lazzaro, B. P. (2017). The genetic architecture of defence as resistance to and tolerance of bacterial infection in Drosophila melanogaster. Mol. Ecol. 26, 1533–1546. doi:10.1111/MEC.14017

PubMed Abstract | CrossRef Full Text | Google Scholar

Huang, W., Campbell, T., Carbone, M. A., Jones, W. E., Unselt, D., Anholt, R. R. H., et al. (2020). Context-dependent genetic architecture of Drosophila life span. PLoS Biol. 18 (3), e3000645. doi:10.1371/JOURNAL.PBIO.3000645

PubMed Abstract | CrossRef Full Text | Google Scholar

Hudry, B., Khadayate, S., and Miguel-Aliaga, I. (2016). The sexual identity of adult intestinal stem cells controls organ size and plasticity. Nature 530 (7590), 344–348. doi:10.1038/nature16953

PubMed Abstract | CrossRef Full Text | Google Scholar

Hyde, R. R. (1913) Inheritance of the length of life in Drosophila ampelophila. bioRxiv. 113–124.

Google Scholar

Jensen, K., McClure, C., Priest, N. K., and Hunt, J. (2015). Sex-specific effects of protein and carbohydrate intake on reproduction but not lifespan in Drosophila melanogaster. Aging Cell 14 (4), 605–615. doi:10.1111/acel.12333

PubMed Abstract | CrossRef Full Text | Google Scholar

Kapahi, P. (2004). Regulation of lifespan in Drosophila by modulation of genes in the TOR signaling pathway. Curr. Biol. CB, 14, 885–890. doi:10.1016/J.CUB.2004.03.059

PubMed Abstract | CrossRef Full Text | Google Scholar

Kelly, C. D., Stoehr, A. M., Nunn, C., Smyth, K. N., and Prokop, Z. M. (2018). Sexual dimorphism in immunity across animals: A meta-analysis. Ecol. Lett. 21 (12), 1885–1894. doi:10.1111/ELE.13164

PubMed Abstract | CrossRef Full Text | Google Scholar

Kierdorf, K. (2020). Muscle function and homeostasis require cytokine inhibition of AKT activity in Drosophila. eLife 9, 51595. doi:10.7554/ELIFE.51595

PubMed Abstract | CrossRef Full Text | Google Scholar

King, E. G. (2012). Genetic dissection of a model complex trait using the Drosophila Synthetic Population Resource. Genome Res. 22, 1558–1566. doi:10.1101/GR.134031.111

PubMed Abstract | CrossRef Full Text | Google Scholar

Kirkwood, T. B. L. (1977). Evolution of ageing. Nature 270 (5635), 301–304. doi:10.1038/270301A0

PubMed Abstract | CrossRef Full Text | Google Scholar

Krejčová, G. (2019). Drosophila macrophages switch to aerobic glycolysis to mount effective antibacterial defense. eLife 8, 50414. doi:10.7554/ELIFE.50414

CrossRef Full Text | Google Scholar

Kurucz, É., Markus, R., Zsamboki, J., Folkl-Medzihradszky, K., Darula, Z., Vilmos, P., et al. (2007). Nimrod, a putative phagocytosis receptor with EGF repeats in Drosophila plasmatocytes, ” Current biology: CB. Curr. Biol. 17 (7), 649–654. doi:10.1016/J.CUB.2007.02.041

PubMed Abstract | CrossRef Full Text | Google Scholar

Kutzer, M. A. M., and Armitage, S. A. O. (2016). The effect of diet and time after bacterial infection on fecundity, resistance, and tolerance in Drosophila melanogaster. Ecol. Evol. 6, 4229. doi:10.1002/ECE3.2185

PubMed Abstract | CrossRef Full Text | Google Scholar

Lazzaro, B. P., Flores, H. A., Lorigan, J. G., and Yourth, C. P. (2008). Genotype-by-Environment interactions and adaptation to local temperature affect immunity and fecundity in Drosophila melanogaster, ” PLoS pathogens. PLoS Pathog. 4 (3), e1000025. doi:10.1371/JOURNAL.PPAT.1000025

PubMed Abstract | CrossRef Full Text | Google Scholar

Lazzaro, B. P., and Little, T. J. (2009). Immunity in a variable worldSeries B, Biological sciences. Philos. Trans. R. Soc. Lond B Biol. Sci. 364, 15–26. doi:10.1098/RSTB.2008.0141

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, H.-Y., Lee, S. H., Lee, J. H., Lee, W. J., and Min, K. J. (2019). The role of commensal microbes in the lifespan of Drosophila melanogaster. Aging 11 (13), 4611–4640. doi:10.18632/aging.102073

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, K. P. (2015). Dietary protein:carbohydrate balance is a critical modulator of lifespan and reproduction in Drosophila melanogaster: A test using a chemically defined diet. J. Insect Physiol. 75, 12–19. doi:10.1016/j.jinsphys.2015.02.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, K. P., Simpson, S. J., Clissold, F. J., Brooks, R., Ballard, J. W. O., Taylor, P. W., et al. (2008). Lifespan and reproduction in Drosophila: New insights from nutritional geometry. Proc. Natl. Acad. Sci. U. S. A. 105 (7), 2498–2503. doi:10.1073/pnas.0710787105

PubMed Abstract | CrossRef Full Text | Google Scholar

Leech, T. (2019). Interactive effects of social environment, age and sex on immune responses in Drosophila melanogaster. J. Evol. Biol. 32, 1082–1092. doi:10.1111/JEB.13509

PubMed Abstract | CrossRef Full Text | Google Scholar

Lemaitre, B., and Hoffmann, J. (2007). The host defense of Drosophila melanogaster. Annu. Rev. Immunol. 25, 697–743. doi:10.1146/ANNUREV.IMMUNOL.25.022106.141615

PubMed Abstract | CrossRef Full Text | Google Scholar

Lemaitre, B., Nicolas, E., Michaut, L., Reichhart, J. M., and Hoffmann, J. A. (1996). The dorsoventral regulatory gene cassette spätzle/Toll/cactus controls the potent antifungal response in Drosophila adults Cell 86 (6), 973–983. doi:10.1016/S0092-8674(00)80172-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Lesser, K. J., Paiusi, I. C., and Leips, J. (2006). Naturally occurring genetic variation in the age-specific immune response of Drosophila melanogaster. Aging Cell 5 (4), 293–295. doi:10.1111/j.1474-9726.2006.00219.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, H., Qi, Y., and Jasper, H. (2016). Preventing age-related decline of gut compartmentalization limits microbiota dysbiosis and extends lifespan. Cell Host Microbe 19 (2), 240–253. doi:10.1016/j.chom.2016.01.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Lind, M. I. (2017). Rapamycin additively extends lifespan in short- and long-lived lines of the nematode Caenorhabditis remanei. Exp. Gerontol. 90, 79–82. doi:10.1016/j.exger.2017.01.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, X., Hodgson, J. J., and Buchon, N. (2017). Drosophila as a model for homeostatic, antibacterial, and antiviral mechanisms in the gut. PLoS Pathog. 13 (5), e1006277. doi:10.1371/journal.ppat.1006277

PubMed Abstract | CrossRef Full Text | Google Scholar

Luckinbill, L. S., and Clare, M. J. (1985). Selection for life span in Drosophila melanogaster, ” Heredity 1985 55:1. Heredity 55 (1), 9–18. doi:10.1038/hdy.1985.66

PubMed Abstract | CrossRef Full Text | Google Scholar

Mackay, T. F. C. (2012). The Drosophila melanogaster genetic reference panel. Nature 482, 173–178. doi:10.1038/nature10811

PubMed Abstract | CrossRef Full Text | Google Scholar

Martin, L. B. (2021). Genes, environments, and phenotypic plasticity in immunology. Trends Immunol. 42, 198–208. doi:10.1016/J.IT.2021.01.002

PubMed Abstract | CrossRef Full Text | Google Scholar

McHugh, K. M., and Burke, M. K. (2022). From microbes to mammals: The experimental evolution of aging and longevity across species. Evolution 76 (4), 692–707. doi:10.1111/evo.14442

PubMed Abstract | CrossRef Full Text | Google Scholar

McKean, K. A., and Nunney, L. (2005). BATEMAN’S principle and immunity: PHENOTYPICALLY plastic reproductive strategies predict changes in immunological sex differences, ” evolution; international journal of organic evolution. Evol. 59 (7), 1510–1517. doi:10.1554/04-657

PubMed Abstract | CrossRef Full Text | Google Scholar

McKean, K. A., Yourth, C. P., Lazzaro, B. P., and Clark, A. G. (2008). The evolutionary costs of immunological maintenance and deployment. BMC Evol. Biol. 8, 76. doi:10.1186/1471-2148-8-76

PubMed Abstract | CrossRef Full Text | Google Scholar

Medzhitov, R., Preston-Hurlburt, P., and Janeway, C. A. (1997). A human homologue of the Drosophila Toll protein signals activation of adaptive immunity. Nature 388 (6640), 394–397. doi:10.1038/41131

PubMed Abstract | CrossRef Full Text | Google Scholar

Metcalf, C. J. E., Roth, O., and Graham, A. L. (2020). Why leveraging sex differences in immune trade-offs may illuminate the evolution of senescence. Funct. Ecol. 34 (1), 129–140. doi:10.1111/1365-2435.13458

PubMed Abstract | CrossRef Full Text | Google Scholar

Meyer, U. A., Zanger, U. M., and Schwab, M. (2013). Omics and drug response. Annu. Rev. Pharmacol. Toxicol. 53, 475–502. doi:10.1146/annurev-pharmtox-010510-100502

PubMed Abstract | CrossRef Full Text | Google Scholar

Millington, J. W., Brownrigg, G. P., Chao, C., Sun, Z., Basner-Collins, P. J., Wat, L. W., et al. (2021). Female-biased upregulation of insulin pathway activity mediates the sex difference in Drosophila body size plasticity. eLife 10, e58341. doi:10.7554/eLife.58341

PubMed Abstract | CrossRef Full Text | Google Scholar

Min, K. J., and Tatar, M. (2018). Unraveling the molecular mechanism of immunosenescence in Drosophila. Int. J. Mol. Sci. 19, 9, doi:10.3390/IJMS19092472

CrossRef Full Text | Google Scholar

Neves, J., and Sousa-Victor, P. (2020). Regulation of inflammation as an anti-aging intervention. FEBS J., 287, 43–52. doi:10.1111/febs.15061

PubMed Abstract | CrossRef Full Text | Google Scholar

Osterwalder, T., Yoon, K. S., White, B. H., and KesHisHian, H. (2001). A conditional tissue-specific transgene expression system using inducible GAL4. Proc. Natl. Acad. Sci. U. S. A. 98 (22), 12596–12601. doi:10.1073/PNAS.221303298

PubMed Abstract | CrossRef Full Text | Google Scholar

Palmer, W. H. (2018). Isolation of a natural DNA virus of Drosophila melanogaster, and characterisation of host resistance and immune responses. Public Libr. Sci. 14, e1007050. doi:10.1371/journal.ppat.1007050

PubMed Abstract | CrossRef Full Text | Google Scholar

Piper, M. D. W., and Partridge, L. (2018). Drosophila as a model for ageing. Biochimica Biophysica Acta (BBA) - Mol. Basis Dis. 1864, 2707–2717. doi:10.1016/J.BBADIS.2017.09.016

CrossRef Full Text | Google Scholar

Ponton, F., Morimoto, J., Robinson, K., Kumar, S. S., Cotter, S. C., Wilson, K., et al. (2020). Macronutrients modulate survival to infection and immunity in Drosophila. J. Anim. Ecol. 89 (2), 460–470. doi:10.1111/1365-2656.13126

PubMed Abstract | CrossRef Full Text | Google Scholar

Rea, I. M. (2018). Age and age-related diseases: Role of inflammation triggers and cytokines. Frontiers in immunology, 9, 586. doi:10.3389/fimmu.2018.00586

PubMed Abstract | CrossRef Full Text | Google Scholar

Regan, J. C. (2020). Dietary restriction and insulin-like signalling pathways as adaptive plasticity: A synthesis and re-evaluationFunct. Ecol. 34, 107–128. doi:10.1111/1365-2435.13418/SUPPINFO

CrossRef Full Text | Google Scholar

Regan, J. C., Khericha, M., Dobson, A. J., Bolukbasi, E., Rattanavirotkul, N., and Partridge, L. (2016). Sex difference in pathology of the ageing gut mediates the greater response of female lifespan to dietary restriction. eLife 5, e10956. doi:10.7554/eLife.10956

PubMed Abstract | CrossRef Full Text | Google Scholar

Regan, J. C., Lu, Y. X., Urena, E., Meilenbrock, R., Catterson, J. H., KiBler, D., et al. (2021). Sexual identity of enterocytes regulates rapamycin-mediated intestinal homeostasis and lifespan extension. bioRxiv. doi:10.1101/2021.10.22.465415

CrossRef Full Text | Google Scholar

Rohde, P. D., Bocker, A., Jensen, C. A. B., Bergstrom, A. L., Madsen, M. I. J., Christensen, S. L., et al. (2021). Genotype and trait specific responses to rapamycin intake in Drosophila melanogaster. Insects 12 (5), 474. doi:10.3390/insects12050474

PubMed Abstract | CrossRef Full Text | Google Scholar

Rose, M. R., and Charlesworth, B. (1981). Genetics of life history in DROSOPHILA MELANOGASTER. II. Exploratory selection experiments. Genetics 97 (1), 187–196. doi:10.1093/GENETICS/97.1.187

PubMed Abstract | CrossRef Full Text | Google Scholar

Rose, M. R. (1991). Evolutionary biology of aging. New York: Oxford University Press.

Google Scholar

Salz, H. K., and Erickson, J. W. (2010). Sex determination in Drosophila: The view from the top. Fly 4 (1), 60–70. doi:10.4161/fly.4.1.11277

PubMed Abstract | CrossRef Full Text | Google Scholar

Sampathkumar, N. K., Bravo, J. I., Chen, Y., Danthi, P. S., Donahue, E. K., Lai, R. W., et al. (2020). Widespread sex dimorphism in aging and age-related diseases. Hum. Genet. 139 (3), 333–356. doi:10.1007/s00439-019-02082-w

PubMed Abstract | CrossRef Full Text | Google Scholar

Sanchez Bosch, P., Makhijani, K., Herboso, L., Gold, K. S., Baginsky, R., Woodcock, K. J., et al. (2019). Adult Drosophila Lack Hematopoiesis but Rely on a Blood Cell Reservoir at the Respiratory Epithelia to Relay Infection Signals to Surrounding Tissues. Dev. cell 51 (6), 787–803.e5. doi:10.1016/j.devcel.2019.10.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Schneider, D. S. (2021). Immunology’s intolerance of disease tolerance. Nat. Rev. Immunol. 21 (10), 624–625. doi:10.1038/s41577-021-00619-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Sciambra, N., and Chtarbanova, S. (2021). The impact of age on response to infection in Drosophila, ” microorganisms. Microorganisms 9 (5), 958. doi:10.3390/MICROORGANISMS9050958

PubMed Abstract | CrossRef Full Text | Google Scholar

Shahrestani, P., Chambers, M., Vandenberg, J., Garcia, K., Malaret, G., Chowdhury, P., et al. (2018). Sexual dimorphism in Drosophila melanogaster survival of Beauveria bassiana infection depends on core immune signaling, ” Scientific reports. Sci. Rep. 8 (1), 12501. doi:10.1038/s41598-018-30527-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Shahrestani, P., King, E., Ramezan, R., Phillips, M., Riddle, M., Thronburg, M., et al. (2021). The molecular architecture of Drosophila melanogaster defense against Beauveria bassiana explored through evolve and resequence and quantitative trait locus mapping. G3 Genes, 11, 324. doi:10.1093/G3JOURNAL/JKAB324

CrossRef Full Text | Google Scholar

Shanley, D. P., and Kirkwood, T. B. L. (2000). Calorie restriction and aging: A life-history analysis, ” evolution; international journal of organic evolution. Evolution 54 (3), 740–750. doi:10.1111/J.0014-3820.2000.TB00076.X

PubMed Abstract | CrossRef Full Text | Google Scholar

Shit, B., Prakash, A., Sarkar, S., Vale, P. F., and Khan, I. (2022). Ageing leads to nonspecific antimicrobial peptide responses in Drosophila melanogaster. bioRxiv. doi:10.1101/2022.06.25.497570

CrossRef Full Text | Google Scholar

Siva-Jothy, J. A., and Vale, P. F. (2021). Dissecting genetic and sex-specific sources of host heterogeneity in pathogen shedding and spread. PLoS Pathog. 17 (1), e1009196. doi:10.1371/journal.ppat.1009196

PubMed Abstract | CrossRef Full Text | Google Scholar

Slack, C., Alic, N., Foley, A., Cabecinha, M., Hoddinott, M. P., and Partridge, L. (2015). The ras-erk-ETS-signaling pathway is a drug target for longevity. Cell 162 (1), 72–83. doi:10.1016/j.cell.2015.06.023

PubMed Abstract | CrossRef Full Text | Google Scholar

Soares, M. P., Teixeira, L., and Moita, L. F. (2017). Disease tolerance and immunity in host protection against infection. Nat. Rev. Immunol. 1717, 283–96. doi:10.1038/nri.2016.136

PubMed Abstract | CrossRef Full Text | Google Scholar

Strong, R., Miller, R. A., Bogue, M., Fernandez, E., Javors, M. A., Libert, S., et al. (2020). Rapamycin-mediated mouse lifespan extension: Late-life dosage regimes with sex-specific effects. Aging Cell 19 (11), e13269. doi:10.1111/acel.13269

PubMed Abstract | CrossRef Full Text | Google Scholar

Tatar, M., KopelmAn, A., Epstein, D., Tu, M. P., Yin, C. M., and Garofalo, R. S. (2001). A mutant Drosophila insulin receptor homolog that extends life-span and impairs neuroendocrine function, ” science. Science 292 (5514), 107–110. doi:10.1126/science.1057987

PubMed Abstract | CrossRef Full Text | Google Scholar

Troha, K., Hyum, J., Lazzaro, B. P., and Buchon, N. (2018). Comparative transcriptomics reveals CrebA as a novel regulator of infection tolerance in D. melanogaster. Public Libr. Sci. 14, e1006847. doi:10.1371/journal.ppat.1006847

PubMed Abstract | CrossRef Full Text | Google Scholar

Unckless, R. L., Rottschaefer, S. M., and Lazzaro, B. P. (2015). The complex contributions of genetics and nutrition to immunity in Drosophila melanogaster. PLoS Genet. 11 (3), e1005030. doi:10.1371/JOURNAL.PGEN.1005030

PubMed Abstract | CrossRef Full Text | Google Scholar

United Nations (2019). World population prospects 2019 highlights. Available at: https://www.un.org/development/desa/publications/world-population-prospects-2019-highlights.html.

Google Scholar

Varma, D., Bülow, M. H., Pesch, Y. Y., Loch, G., and Hoch, M. (2014). Forkhead, a new cross regulator of metabolism and innate immunity downstream of TOR in Drosophila. J. insect physiology 69, 80–88. doi:10.1016/J.JINSPHYS.2014.04.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Vellai, T., Takacs-Vellai, K., Zhang, Y., Kovacs, A. L., Orosz, L., and Muller, F. (2003). Influence of TOR kinase on lifespan in C. elegans, ” Nature. Nature 426 (6967), 620–620. doi:10.1038/426620a

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, J. B., Lu, H. L., and St. Leger, R. J. (2017). The genetic basis for variation in resistance to infection in the Drosophila melanogaster genetic reference panel. Public Libr. Sci. 13 (3), e1006260. doi:10.1371/JOURNAL.PPAT.1006260

PubMed Abstract | CrossRef Full Text | Google Scholar

Williams, G. C. (1957). Pleiotropy, natural selection, and the evolution of senescence, ” evolution; international journal of organic evolution. Evolution 11 (4), 398. doi:10.2307/2406060

CrossRef Full Text | Google Scholar

Xirocostas, Z. A., Everingham, S. E., and Moles, A. T. (2020). The sex with the reduced sex chromosome dies earlier: A comparison across the tree of life. Biol. Lett. 16 (3), 20190867. doi:10.1098/RSBL.2019.0867

PubMed Abstract | CrossRef Full Text | Google Scholar

Zanco, B., Mirth, C. K., Sgro, C. M., and Piper, M. D. (2021). A dietary sterol trade-off determines lifespan responses to dietary restriction in Drosophila melanogaster females. eLife 10, e62335. doi:10.7554/eLife.62335

PubMed Abstract | CrossRef Full Text | Google Scholar

Zerofsky, M., Harel, E., Silverman, N., and Tatar, M. (2005). Aging of the innate immune response in Drosophila melanogaster Aging Cell 4 (2), 103–108. doi:10.1111/j.1474-9728.2005.00147.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: Drosophila, immunosenescence, immunity, ageing, natural variation

Citation: Corbally M-K and Regan JC (2022) Fly immunity comes of age: The utility of Drosophila as a model for studying variation in immunosenescence. Front. Aging 3:1016962. doi: 10.3389/fragi.2022.1016962

Received: 11 August 2022; Accepted: 20 September 2022;
Published: 04 October 2022.

Edited by:

Susan Broughton, Lancaster University, United Kingdom

Reviewed by:

Myriam Richaud, UMR5247 Institut des Biomolécules Max Mousseron (IBMM), France

Copyright © 2022 Corbally and Regan. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Jennifer C. Regan, jenny.regan@ed.ac.uk

Download