Skip to main content

MINI REVIEW article

Front. Nutr., 11 April 2024
Sec. Food Chemistry
Volume 11 - 2024 | https://doi.org/10.3389/fnut.2024.1394697

Tastant-receptor interactions: insights from the fruit fly

Christian Arntsen Jacqueline Guillemin Kayla Audette Molly Stanley*
  • Department of Biology, University of Vermont, Burlington, VT, United States

Across species, taste provides important chemical information about potential food sources and the surrounding environment. As details about the chemicals and receptors responsible for gustation are discovered, a complex view of the taste system is emerging with significant contributions from research using the fruit fly, Drosophila melanogaster, as a model organism. In this brief review, we summarize recent advances in Drosophila gustation and their relevance to taste research more broadly. Our goal is to highlight the molecular mechanisms underlying the first step of gustatory circuits: ligand-receptor interactions in primary taste cells. After an introduction to the Drosophila taste system and how it encodes the canonical taste modalities sweet, bitter, and salty, we describe recent insights into the complex nature of carboxylic acid and amino acid detection in the context of sour and umami taste, respectively. Our analysis extends to non-canonical taste modalities including metals, fatty acids, and bacterial components, and highlights unexpected receptors and signaling pathways that have recently been identified in Drosophila taste cells. Comparing the intricate molecular and cellular underpinnings of how ligands are detected in vivo in fruit flies reveals both specific and promiscuous receptor selectivity for taste encoding. Throughout this review, we compare and contextualize these Drosophila findings with mammalian research to not only emphasize the conservation of these chemosensory systems, but to demonstrate the power of this model organism in elucidating the neurobiology of taste and feeding.

Introduction

The chemical sense of taste allows animals to evaluate their food options to encourage the consumption of beneficial nutrients and avoidance of potential toxins. Since gustation links the environment to nutrition and fitness, it is not surprising that this sense is well-conserved across a wide range of animals, from humans to fruit flies (1). The concept that certain chemicals elicit distinct taste perceptions can be traced back to the earliest philosophers, but a clear understanding of the molecular and cellular basis of taste only started to emerge in the early 2000s. Over the last two decades, there has been extensive research into identifying the receptors responsible for the “five basic tastes”: sweet, bitter, salty, sour, and umami (2). Many details of these canonical taste modalities are well-established in both mammalian and non-mammalian model organisms, including the fruit fly, Drosophila melanogaster (1, 35). Drosophila is a powerful model organism in neurobiology research that has continued to advance our understanding of gustation due to the ability to record taste cell activity in vivo from a single neuron or a complete set of specific taste cells (68). Readily available genetic tools also allow for investigation into the role of taste receptors in cellular physiology and chemosensory behaviors (911). This review introduces the Drosophila taste system and describes recent insights into novel tastant-receptor interactions for both canonical and non-canonical taste modalities with comparisons to mammalian gustation.

The fruit fly taste system

In both mammals and Drosophila, primary chemosensory cells initiate taste sensation by evaluating a food source’s chemical properties. The mammalian gustatory system uses taste receptor cells (TRCs), modified epithelial cells found in taste buds throughout the oral cavity. TRCs detect chemicals and relay this information to afferent gustatory nerves (1), but the Drosophila peripheral nervous system directly detects tastants via gustatory receptor neurons (GRNs) (12). GRNs are distributed throughout the fly body, but the highest concentration of taste cells involved in feeding is located in the labellum, the Drosophila tongue homolog (5). Labellar GRNs express taste receptors that allow for the rapid identification of chemicals, promoting selectivity for compounds that represent specific taste modalities (13, 14), akin to lingual taste cells in mammals.

GRNs in the fruit fly labellum have been categorized into five groups based on their distinct receptor profiles and taste modality responsiveness: “sweet,” “bitter,” “water,” “salty,” and “IR94e” (15). These five GRN classes can be consistently mapped on a fly’s labellum across ~62 gustatory sensilla that are classified by size, each containing two or four GRNs (10, 16, 17) (Figure 1A). GRN axons project to the sub-esophageal zone in the brain (17, 18), where arborizations of both GRNs and motor neurons generate local circuits for taste-induced behavioral responses (19). The Drosophila whole-brain connectome (2023) allows neural circuits to be traced from tastant-receptor activation through behavioral output to enhance our understanding of how taste information is encoded and modulated (2428).

Figure 1
www.frontiersin.org

Figure 1. Neural encoding of novel tastant-receptor interactions in Drosophila. (A) Fruit flies will explore their chemosensory environment by extending their proboscis and interacting with tastants via sensillum on the labellum. (A inset) Three types of sensilla (S = short, I = intermediate, L = long) each contain two or four gustatory receptor neurons (GRNs) from five different cell types (color depicts each class, grouped by modality and receptor expression). Mechanosensory neurons are depicted in black. (B) Each GRN has a diverse array of chemosensory receptors (colored to match the GRN that houses it). Recently investigated tastants are depicted (gray boxes) with their specific receptors (colored arrows). * indicates “tuning” IRs that generally work with the ionotropic co-receptors. (*) indicates receptors that work alongside ionotropic co-receptors for only some of the taste modalities depicted. Tastant-receptor pairings that require intracellular pathways are depicted (gray arrows). Created with Biorender.com.

Gustatory processing commonly starts with two main classes of taste receptors in Drosophila: gustatory receptors (GRs) and ionotropic receptors (IRs). GRs consist of seven transmembrane domains, an intracellular N-terminus, and an extracellular C-terminus (2932). Earlier studies that disrupted G-protein subunits in GR-expressing cells found a reduction in taste responses (33, 34), but recent research elucidating the structure of two sugar GRs has determined that they form tetrameric ligand-gated cation channels with peripheral ligand binding sites and a single central pore (35). The other class of Drosophila taste receptors, IRs, share structural similarities with synaptic, glutamate-gated ion channels: 3 transmembrane domains and a 2-lobed extracellular binding domain (3638). IRs form heteromeric receptor complexes comprised of both co-receptors and “tuning” receptors that function as ligand-gated ion channels (39). In contrast to IRs and GRs, mammalian T1Rs and T2Rs are G-protein coupled receptors (GPCRs) (4042), yet the repertoire of tastants acting via these receptors and their impact on behavior is remarkably conserved (1).

Sweet, bitter, and salty taste

Direct activation of Drosophila “sweet” GRNs leads to appetitive feeding behaviors whereas “bitter” GRN activation produces avoidance (15, 19, 43), consistent with mammalian studies on these canonical tastes (1, 44, 45). Drosophila sugar receptors formed from 9 GR genes (sugar GRs) (4651) detect mono- and disaccharides along with artificial sweeteners and molecules perceived as sweet to humans (5254). The remaining 30+ GR genes form receptors in “bitter” GRNs (bitter GRs), detecting a range of bitter compounds (e.g., caffeine, lobeline, denatonium, and quinine) (5558). Recent work in Drosophila has identified two non-canonical bitter signaling pathways for the detection of specific ligands, using rhodopsins and a peptidoglycan recognition protein (PGRP), that open new avenues for taste transduction (Table 1; Figure 1B). Rhodopsin GPCRs are typically light-sensitive with an opsin protein and retinal chromophore, but three rhodopsins (Rh1, Rh4, and Rh7) were found to function as taste receptors that do not require light or retinal (59). These rhodopsins detect aristolochic acid and activate “bitter” GRNs at particularly low concentrations through a phospholipase C (PLC) signaling cascade that involves TRPA1 (59). Mouse taste buds express some opsin RNA (75), suggesting these channels may have a conserved role in chemosensation. The other non-canonical pathway involves PGRPs, pattern recognition receptors traditionally involved in the immune response to pathogens. TRPA1 and canonical bitter GRs (Gr33a, Gr66a) were previously implicated in the detection of bacterial components (76, 77), but the newly described PGRP (PGRP-LB) expressed in the labellum specifically detects bacterial peptidoglycans. Unexpectedly, this receptor uses nuclear factor-κB (NF-κB)/immune deficiency (IMD)-dependent signaling to activate “bitter” GRNs (60). An interest in the role of oral taste receptors in microbial detection has emerged in mammalian research (78, 79), and this recent work in fruit flies highlights an unexpected role for NF-κB/IMD signaling in taste cells that impacts feeding choices (60).

Table 1
www.frontiersin.org

Table 1. Recently described tastants and their receptors in the Drosophila labellum.

Recent advances in salt taste have revealed a complex taste transduction system that allows for concentration-dependent salt feeding in both mammals and fruit flies (80). A set of “high salt” or “salty” GRNs in the Drosophila labellum are specifically activated by high concentrations of various salt ions (15, 81, 82). Salt also activates other GRNs (“sweet,” “bitter,” and “IR94e”) while inhibiting “water” GRNs, producing a combinatorial code that can lead to flexible behaviors (15). Salt taste research highlights the role of IRs that use the broadly expressed co-receptors, IR25a and IR76b, plus a narrowly expressed “tuning” IR to form functional receptors that detect specific salt ions (15, 8284). Canonical salt taste centers around NaCl and occasionally other mono- or divalent ions (80), but recent research has shifted focus to identify the taste mechanisms for other ions.

Metal taste

Metals, including divalent and trivalent salt ions, have complex taste profiles (8587) that have garnered increasing attention due to their accumulation in soil, crops, and foods from human activities (88, 89). Recent studies established that the human bitter taste receptor TAS2R7 acts as a metal cation receptor for detecting zinc and copper (90), yet this can only be demonstrated in vitro. Fruit flies avoid consuming metals and in vivo quantifications of neuronal activity reveal that metal ions activate taste cells through multiple receptors (Table 1; Figure 1B).

In Drosophila, some metal ions require only bitter GRs (Cu2+, Ag+) or IRs (Mn2+, Ni2+, Cd2+) for detection, while others require both (Zn2+, Co2+). Interestingly, cellular responses to iron involve both receptor types or solely IRs depending if it is in the Fe2+ or Fe3+ form, respectively (61). Cadmium sensitivity requires an IR complex in two types of GRNS: co-receptors (IR25a, IR76b) plus IR7a in “bitter” GRNs and the same co-receptors plus IR47a in “salty” GRNs (63). A recent brief report found an additional “tuning” receptor, IR56b, to be necessary for zinc avoidance (62), however, this receptor complex detects NaCl in “sweet” GRNs for attraction (83), so this is an unexpected result. Overall, this research in flies provides clear evidence that a range of individual metal ions have specific taste detection mechanisms. As metal contamination continues to rise, understanding gustatory pathways for metal ligands will become increasingly important across animals with relevance to environmental health and food safety.

Sour taste

pH is an important indicator of food quality and sour taste describes the gustatory detection of acids. Recently, the Otop1 proton channel was identified as the “sour receptor” in mammals (91, 92), and a homolog, OtopLa, is expressed in fruit fly GRNs in the labellum (93, 94). While the discovery of Otop channels was an important breakthrough for sour taste, different acids have distinct taste qualities even at the same pH, suggesting there is more to sour taste than pH alone (95, 96).

Like humans, fruit flies show dose-dependent attraction or aversion to certain carboxylic acids (67). Weak organic acids, such as acetic acid, may have the ability to cross the membrane of taste cells to impact transduction by altering intracellular pH, but through unknown mechanisms (97, 98). In Drosophila, attractive concentrations of specific organic acids—acetic, lactic, glycolic, and citric—require taste receptors for the activation of “sweet” GRNs (6466). At least one broadly expressed co-receptor, IR25a, is involved, along with sugar GRs for detecting organic acids (6466) (Table 1; Figure 1B). Even at the same pH, these acids differentially activate Drosophila taste cells in vivo, indicating diverse receptor binding and/or abilities to cross cell membranes. Attempts to distinguish between the detection of pH and anion species show that IRs are largely involved with anion detection, whereas the sugar GRs are responsive to the change in pH (66). Ascorbic acid (Vitamin C), a distinct but related acidic compound, was also found to activate “sweet” GRNs through similar mechanisms (65). High concentrations of carboxylic acids are aversive (99), and IR7a in “bitter” GRNs is specifically required for acetic acid avoidance, without the need for IR co-receptors (67). While the cooperative role of OtopLa channels along with these receptors remains unclear, these findings underscore the dual activation of taste cells by acids through both receptors and proton influx.

Alkaline taste

Since pH influences food quality, the ability to detect both basic and acidic pH levels would be advantageous. Previous mammalian studies on basic pH sensation focused on somatosensation (100), but humans show alkaline sensitivity on the tip of the tongue (101) and a recent study in rats found that sodium carbonate (Na2CO3) solutions activate taste nerves significantly more than Na+ alone. However, alkaline taste has not been well described. A recent study in Drosophila established the existence of alkaline taste, and identified a novel receptor required for the detection of basic solutions (70) (Table 1; Figure 1B). Alkaliphile (Alka) is a Cl channel gated by high pH that is necessary for alkaline taste (70). The Alka receptor is expressed in a subset of the “bitter” GRNs (~21%), but it is currently unclear what other cell types may express this receptor. Regardless, this study in flies establishes a novel tastant-receptor interaction for alkaline taste that may be relevant to mammals. Interestingly, the Otop1 proton channel for sour taste was recently found to be a candidate alkaline receptor in vitro (102), indicating a need for future comparative studies on basic and acidic pH detection mechanisms.

Umami (amino acid) taste

Protein feeding is coupled with the chemical detection of amino acids. Umami taste is a specific savory sensation, usually associated with monosodium L-glutamate (MSG), an amino acid often found in foods at higher concentrations (103105). The mammalian GPCR complex consisting of T1R1 + T1R3 is referred to as the “umami receptor” (45) and has a high sensitivity to glutamate in humans (106). In most vertebrates, this receptor is broadly responsive to amino acid ligands and amino acids can also activate sugar taste receptors, bitter taste receptors, or act through metabotropic glutamate receptors in multiple cell types (107114). This combinatorial coding likely occurs in response to individual amino acids in a dose-dependent manner. Through in vitro assays, mammalian bitter receptors display dose-dependent activation by amino acids, however, some inconclusive results are attributed to the possibility of endogenous amino acid receptors in the cell line used for these experiments (109). In vivo studies in Drosophila circumvent these concerns and allow for a deeper understanding of the combinatorial coding for amino acid taste (Table 1).

Fruit flies require and consume amino acids based on internal state, such as mating status or nutritional deficiency (115). The IR co-receptors (IR76b and IR25a) are necessary for detecting most of the proteinogenic amino acids at various concentrations (68, 69, 116), and the “tuning” receptors identified for amino acid sensation to date are IR51b and IR94e. IR51b is a bitter cell-specific receptor that detects high concentrations of arginine, valine, leucine, tryptophan, isoleucine, lysine, and proline (68). IR94e receptors are integral for the detection of glutamate in various forms, and this “tuning” receptor is expressed in a newly described set of taste cells that induce mild feeding aversion (27, 69). A thorough description of the combinatorial coding for low concentrations (25 mM) of arginine reveals that “sweet” GRNs are activated through both sugar GRs (Gr5a, Gr61a, and Gr64f) and IR co-receptors (68) (Figure 1B). The overlap between sugar-sensing and amino acid-sensing resembles a pattern found in mammals (117).

A feature of the mammalian “umami receptor” is enhancement by purine-5′-nucleotides (IMP and GMP) (111, 114, 118), but this feature is not known to occur in fruit flies. Additionally, while no metabotropic glutamate receptors have been identified in fruit fly amino acid taste, the IRs are ancestrally related to ionotropic glutamate receptors (3638, 119), suggesting a conserved use of glutamate receptors in chemosensation (107, 113, 120). In Drosophila, another intriguing element is that an odorant binding protein (OBP19b) secreted from nearby cells can bind certain amino acids to impact their detection by taste cells (121), but it is unclear how conserved this mechanism may be. Despite some differences from the mammalian system, the Drosophila model offers a way to study dose-dependent encoding of individual or groups of amino acids to better understand this canonical yet complex taste modality.

Fatty acid taste

Fatty acids are highly energetic essential nutrients that are attractive to both mammals and Drosophila (71, 122124). Initially, fat palatability was thought to be driven by texture and olfaction (125), but more recent research has highlighted the importance of gustation (126128). In mice, CD36 is a fatty acid transporter expressed in taste buds that contributes to fatty acid preferences (129), and two GPCRs (GPR40 and GPR120) appear to function as lingual fat receptors (130). Although Drosophila homologs have not been discovered, GRNs in the labellum do detect fatty acids (71). Similar to carboxylic and amino acids, the cellular and behavioral responses to fatty acids in flies depend on concentration.

At low concentrations (~0.1%), hexanoic acid elicits appetitive responses in Drosophila, while at high concentrations (~1–2%), it prompts aversion (74). Hexanoic acid attraction is driven by “sweet” GRNs, requiring both IR56d and Gr64d (7274, 131, 132). Aversion to hexanoic acid is controlled by “bitter” GRNs via three bitter receptors: Gr32a, Gr33a, and Gr66a (74) (Table 1). Recent work has also demonstrated that the fly gustatory system can distinguish between different classes of fatty acids based on chain length (73). While all classes of fatty acids require the IR co-receptors (IR25a and IR76b) for detection, medium-chain fatty acid (MCFA) taste requires “sweet” GRNs and the IR56d receptor, whereas short-chain (SCFA) and long-chain (LCFA) fatty acid taste does not (73). These findings indicate that IR56d is selective for MCFAs, while the co-receptors may function more broadly. However, a recent study questioned the involvement of IR25a and IR76b in the labellar response to the MCFA hexanoic acid (74). The molecular and cellular underpinnings of SCFA/LCFA detection and fatty acid discrimination remain unclear, but these complexities reflect a nuanced fatty acid taste encoding system that is sensitive to both concentration and subtle variations in molecular structure.

MCFA taste also requires intracellular signaling, as flies with a mutant norpA, a Drosophila homolog for PLC, have disrupted MCFA detection (71) (Figure 1B). Whether or not PLC signaling is necessary for SCFA and LCFA sensation is unknown. Furthermore, one study showed that the sugar GR, Gr64e, is an essential component of MCFA signal transduction, unexpectedly serving as a downstream component in the PLC pathway within “sweet” GRNs (133). Notably, a recent investigation found that Gr64e mutation did not affect electrophysiological responses to the MCFA hexanoic acid (74). Despite this discrepancy, activation of a secondary receptor via PLC mimics the mammalian fatty acid signaling cascade. Mice lacking PLC or TRPM5, a downstream receptor in the PLC cascade, lose their taste preference for fatty acids (134). Collectively, these results imply that PLC-mediated intracellular mechanisms underpin fatty acid gustation in both Drosophila and mammals, despite mammalian research primarily focusing on LCFAs which remain attractive at higher concentrations (130). Drosophila fatty acid taste emphasizes the conserved nature of macronutrient taste encoding and may prove valuable for informing future fat perception research to uncover more about this non-canonical taste modality that has many health implications.

Discussion

Recent advances in gustation research using Drosophila melanogaster as a model organism have revealed several unexpected ligand-receptor interactions within the taste system that play crucial roles in chemosensation and behavior. The discovery of two novel receptor signaling types in bitter cells, through non-canonical rhodopsin and immune signaling, has revealed unexpected transducers for contact chemical cues. Moreover, the fly gustatory system contains a markedly complex set of receptors to detect specific metals, which may become increasingly relevant in this Anthropocene Epoch. The identification of receptors for carboxylic acid anions suggests a mechanism for sour taste that extends beyond proton detection, while a novel receptor for alkaline solutions highlights the role of gustation in discerning a broader pH spectrum. The ability to study intact taste cells in awake flies has provided key insights into the concentration-dependent nature of ligand detection across multiple receptors and cell types for carboxylic, amino, and fatty acids that imply combinatorial taste coding mechanisms to specific molecules. Future work can apply these insights to continue understanding the repertoire of tastant-receptor interactions behind basic, canonical tastes and emerging, non-canonical taste modalities.

Author contributions

CA: Writing – review & editing, Writing – original draft, Visualization, Conceptualization. JG: Writing – review & editing, Writing – original draft, Visualization, Conceptualization. KA: Writing – review & editing, Visualization. MS: Writing – review & editing, Writing – original draft, Visualization, Supervision, Funding acquisition, Conceptualization.

Funding

The author(s) declare that financial support was received for the research, authorship, and/or publication of this article. The University of Vermont new lab startup funds were used to support this publication.

Acknowledgments

BioRender was used to generate Figure 1.

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

1. Yarmolinsky, DA, Zuker, CS, and Ryba, NJ. Common sense about taste: from mammals to insects. Cell. (2009) 139:234–44. doi: 10.1016/j.cell.2009.10.001

PubMed Abstract | Crossref Full Text | Google Scholar

2. Beauchamp, GK. Basic Taste: A Perceptual Concept. J Agric Food Chem. (2019) 67:13860–9. doi: 10.1021/acs.jafc.9b03542

PubMed Abstract | Crossref Full Text | Google Scholar

3. Chandrashekar, J, Hoon, MA, Ryba, NJP, and Zuker, CS. The receptors and cells for mammalian taste. Nature. (2006) 444:288–94. doi: 10.1038/nature05401

Crossref Full Text | Google Scholar

4. Montell, C. A taste of the Drosophila gustatory receptors. Curr Opin Neurobiol. (2009) 19:345–53. doi: 10.1016/j.conb.2009.07.001

PubMed Abstract | Crossref Full Text | Google Scholar

5. Scott, K. Gustatory processing in Drosophila melanogaster. Annu Rev Entomol. (2018) 63:15–30. doi: 10.1146/annurev-ento-020117-043331

Crossref Full Text | Google Scholar

6. Delventhal, R, Kiely, A, and Carlson, JR. Electrophysiological recording from Drosophila labellar taste sensilla. J Vis Exp. (2014) 2014:e51355. doi: 10.3791/51355

PubMed Abstract | Crossref Full Text | Google Scholar

7. Dweck, HKM, and Carlson, JR. Diverse mechanisms of taste coding in Drosophila. Science. Advances. (2023) 9:eadj7032. doi: 10.1126/sciadv.adj7032

Crossref Full Text | Google Scholar

8. Marella, S, Mann, K, and Scott, K. Dopaminergic modulation of sucrose acceptance behavior in Drosophila. Neuron. (2012) 73:941–50. doi: 10.1016/j.neuron.2011.12.032

PubMed Abstract | Crossref Full Text | Google Scholar

9. Chen, Y-CD, and Dahanukar, A. Recent advances in the genetic basis of taste detection in Drosophila. Cell Mol Life Sci. (2020) 77:1087–101. doi: 10.1007/s00018-019-03320-0

PubMed Abstract | Crossref Full Text | Google Scholar

10. Freeman, EG, and Dahanukar, A. Molecular neurobiology of Drosophila taste. Curr Opin Neurobiol. (2015) 34:140–8. doi: 10.1016/j.conb.2015.06.001

PubMed Abstract | Crossref Full Text | Google Scholar

11. Montell, C. Drosophila sensory receptors-a set of molecular Swiss Army knives. Genetics. (2021) 217:1–34. doi: 10.1093/genetics/iyaa011

PubMed Abstract | Crossref Full Text | Google Scholar

12. Singh, RN. Neurobiology of the gustatory systems of Drosophila and some terrestrial insects. Microsc Res Tech. (1997) 39:547–63. doi: 10.1002/(SICI)1097-0029(19971215)39:6<547::AID-JEMT7>3.0.CO;2-A

PubMed Abstract | Crossref Full Text | Google Scholar

13. Fujishiro, N, Kijima, H, and Morita, H. Impulse frequency and action potential amplitude in labellar chemosensory neurones of Drosophila melanogaster. J Insect Physiol. (1984) 30:317–25. doi: 10.1016/0022-1910(84)90133-1

Crossref Full Text | Google Scholar

14. Ishimoto, H, and Tanimura, T. Molecular neurophysiology of taste in Drosophila. Cell Mol Life Sci. (2004) 61:10–8. doi: 10.1007/s00018-003-3182-9

PubMed Abstract | Crossref Full Text | Google Scholar

15. Jaeger, AH, Stanley, M, Weiss, ZF, Musso, PY, Chan, RC, Zhang, H, et al. A complex peripheral code for salt taste in Drosophila. eLife. (2018) 7:e37167. doi: 10.7554/eLife.37167

PubMed Abstract | Crossref Full Text | Google Scholar

16. Falk, R, Bleiser-Avivi, N, and Atidia, J. Labellar taste organs of Drosophila melanogaster. J Morphol. (1976) 150:327–41. doi: 10.1002/jmor.1051500206

PubMed Abstract | Crossref Full Text | Google Scholar

17. Stocker, RF. The organization of the chemosensory system in Drosophila melanogaster: a review. Cell Tissue Res. (1994) 275:3–26. doi: 10.1007/BF00305372

PubMed Abstract | Crossref Full Text | Google Scholar

18. Stocker, R, and Schorderet, M. Cobalt filling of sensory projections from internal and external mouthparts in Drosophila. Cell Tissue Res. (1981) 216:513–23. doi: 10.1007/BF00238648

PubMed Abstract | Crossref Full Text | Google Scholar

19. Gordon, MD, and Scott, K. Motor control in a Drosophila taste circuit. Neuron. (2009) 61:373–84. doi: 10.1016/j.neuron.2008.12.033

PubMed Abstract | Crossref Full Text | Google Scholar

20. Dorkenwald, S, Matsliah, A, Sterling, AR, Schlegel, P, Yu, SC, Mckellar, CE, et al. Neuronal wiring diagram of an adult brain. bioRxiv. (2023) 11:2023.06.27.546656. doi: 10.1101/2023.06.27.546656

PubMed Abstract | Crossref Full Text | Google Scholar

21. Dorkenwald, S, Mckellar, CE, Macrina, T, Kemnitz, N, Lee, K, Lu, R, et al. FlyWire: online community for whole-brain connectomics. Nat Methods. (2022) 19:119–28. doi: 10.1038/s41592-021-01330-0

PubMed Abstract | Crossref Full Text | Google Scholar

22. Scheffer, LK, Xu, CS, Januszewski, M, Lu, Z, Takemura, S-Y, Hayworth, KJ, et al. A connectome and analysis of the adult Drosophila central brain. eLife. (2020) 9:e57443. doi: 10.7554/eLife.57443

PubMed Abstract | Crossref Full Text | Google Scholar

23. Zheng, Z, Lauritzen, JS, Perlman, E, Robinson, CG, Nichols, M, Milkie, D, et al. A complete Electron microscopy volume of the brain of adult Drosophila melanogaster. Cell. (2018) 174:730–743.e22. doi: 10.1016/j.cell.2018.06.019

PubMed Abstract | Crossref Full Text | Google Scholar

24. Deere, JU, Sarkissian, AA, Yang, M, Uttley, HA, Martinez Santana, N, Nguyen, L, et al. Selective integration of diverse taste inputs within a single taste modality. eLife. (2023) 12:e84856. doi: 10.7554/eLife.84856

PubMed Abstract | Crossref Full Text | Google Scholar

25. Jacobs, RV, Wang, CX, Lozada-Perdomo, FV, Nguyen, L, Deere, JU, Uttley, HA, et al. Overlap and divergence of neural circuits mediating distinct behavioral responses to sugar. bioRxiv. (2023). doi: 10.1101/2023.10.01.560401

Crossref Full Text | Google Scholar

26. Shiu, PK, Sterne, GR, Engert, S, Dickson, BJ, and Scott, K. Taste quality and hunger interactions in a feeding sensorimotor circuit. eLife. (2022) 11:e79887. doi: 10.7554/eLife.79887

PubMed Abstract | Crossref Full Text | Google Scholar

27. Shiu, PK, Sterne, GR, Spiller, N, Franconville, R, Sandoval, A, Zhou, J, et al. A leaky integrate-and-fire computational model based on the connectome of the entire adult Drosophila brain reveals insights into sensorimotor processing. bioRxiv. (2023). doi: 10.1101/2023.05.02.539144

Crossref Full Text | Google Scholar

28. Sterne, GR, Otsuna, H, Dickson, BJ, and Scott, K. Classification and genetic targeting of cell types in the primary taste and premotor center of the adult Drosophila brain. eLife. (2021) 10:e71679. doi: 10.7554/eLife.71679

PubMed Abstract | Crossref Full Text | Google Scholar

29. Clyne, PJ, Warr, CG, and Carlson, JR. Candidate taste receptors in Drosophila. Science. (2000) 287:1830–4. doi: 10.1126/science.287.5459.1830

PubMed Abstract | Crossref Full Text | Google Scholar

30. Dunipace, L, Meister, S, Mcnealy, C, and Amrein, H. Spatially restricted expression of candidate taste receptors in the Drosophila gustatory system. Curr Biol. (2001) 11:822–35. doi: 10.1016/S0960-9822(01)00258-5

PubMed Abstract | Crossref Full Text | Google Scholar

31. Robertson, HM, Warr, CG, and Carlson, JR. Molecular evolution of the insect chemoreceptor gene superfamily in Drosophila melanogaster. Proc Natl Acad Sci. (2003) 100:14537–42. doi: 10.1073/pnas.2335847100

PubMed Abstract | Crossref Full Text | Google Scholar

32. Scott, K, Brady, R, Cravchik, A, Morozov, P, Rzhetsky, A, Zuker, C, et al. A chemosensory gene family encoding candidate gustatory and olfactory receptors in Drosophila. Cell. (2001) 104:661–73. doi: 10.1016/S0092-8674(01)00263-X

PubMed Abstract | Crossref Full Text | Google Scholar

33. Bredendiek, N, Hütte, J, Steingräber, A, Hatt, H, Gisselmann, G, and Neuhaus, EM. Goα is involved in sugar perception in Drosophila. Chem Senses. (2010) 36:69–81. doi: 10.1093/chemse/bjq100

Crossref Full Text | Google Scholar

34. Ueno, K, Kohatsu, S, Clay, C, Forte, M, Isono, K, and Kidokoro, Y. Gsα is involved in sugar perception in Drosophila melanogaster. J Neurosci. (2006) 26:6143–52. doi: 10.1523/JNEUROSCI.0857-06.2006

PubMed Abstract | Crossref Full Text | Google Scholar

35. Ma, D, Hu, M, Yang, X, Liu, Q, Ye, F, Cai, W, et al. Structural basis for sugar perception by Drosophila gustatory receptors. Science. (2024) 383:eadj2609. doi: 10.1126/science.adj2609

Crossref Full Text | Google Scholar

36. Benton, R, Vannice, KS, Gomez-Diaz, C, and Vosshall, LB. Variant ionotropic glutamate receptors as chemosensory receptors in Drosophila. Cell. (2009) 136:149–62. doi: 10.1016/j.cell.2008.12.001

PubMed Abstract | Crossref Full Text | Google Scholar

37. Croset, V, Rytz, R, Cummins, SF, Budd, A, Brawand, D, Kaessmann, H, et al. Ancient protostome origin of chemosensory ionotropic glutamate receptors and the evolution of insect taste and olfaction. PLoS Genet. (2010) 6:e1001064. doi: 10.1371/journal.pgen.1001064

PubMed Abstract | Crossref Full Text | Google Scholar

38. Rytz, R, Croset, V, and Benton, R. Ionotropic receptors (IRs): chemosensory ionotropic glutamate receptors in Drosophila and beyond. Insect Biochem Mol Biol. (2013) 43:888–97. doi: 10.1016/j.ibmb.2013.02.007

PubMed Abstract | Crossref Full Text | Google Scholar

39. Abuin, L, Bargeton, B, Ulbrich, MH, Isacoff, EY, Kellenberger, S, and Benton, R. Functional architecture of olfactory ionotropic glutamate receptors. Neuron. (2011) 69:44–60. doi: 10.1016/j.neuron.2010.11.042

PubMed Abstract | Crossref Full Text | Google Scholar

40. Andres-Barquin, PJ, and Conte, C. Molecular basis of bitter taste: the T2R family of G protein-coupled receptors. Cell Biochem Biophys. (2004) 41:099–112. doi: 10.1385/CBB:41:1:099

PubMed Abstract | Crossref Full Text | Google Scholar

41. Sainz, E, Cavenagh, MM, Lopezjimenez, ND, Gutierrez, JC, Battey, JF, Northup, JK, et al. The G-protein coupling properties of the human sweet and amino acid taste receptors. Dev Neurobiol. (2007) 67:948–59. doi: 10.1002/dneu.20403

PubMed Abstract | Crossref Full Text | Google Scholar

42. Wong, GT, Gannon, KS, and Margolskee, RF. Transduction of bitter and sweet taste by gustducin. Nature. (1996) 381:796–800. doi: 10.1038/381796a0

Crossref Full Text | Google Scholar

43. Marella, S, Fischler, W, Kong, P, Asgarian, S, Rueckert, E, and Scott, K. Imaging taste responses in the fly brain reveals a functional map of taste category and behavior. Neuron. (2006) 49:285–95. doi: 10.1016/j.neuron.2005.11.037

PubMed Abstract | Crossref Full Text | Google Scholar

44. Mueller, KL, Hoon, MA, Erlenbach, I, Chandrashekar, J, Zuker, CS, and Ryba, NJ. The receptors and coding logic for bitter taste. Nature. (2005) 434:225–9. doi: 10.1038/nature03352

PubMed Abstract | Crossref Full Text | Google Scholar

45. Zhao, GQ, Zhang, Y, Hoon, MA, Chandrashekar, J, Erlenbach, I, Ryba, NJ, et al. The receptors for mammalian sweet and umami taste. Cell. (2003) 115:255–66. doi: 10.1016/S0092-8674(03)00844-4

Crossref Full Text | Google Scholar

46. Dahanukar, A, Lei, Y-T, Kwon, JY, and Carlson, JR. Two gr genes underlie sugar reception in Drosophila. Neuron. (2007) 56:503–16. doi: 10.1016/j.neuron.2007.10.024

PubMed Abstract | Crossref Full Text | Google Scholar

47. Fujii, S, Yavuz, A, Slone, J, Jagge, C, Song, X, and Amrein, H. Drosophila sugar receptors in sweet taste perception, olfaction, and internal nutrient sensing. Curr Biol. (2015) 25:621–7. doi: 10.1016/j.cub.2014.12.058

PubMed Abstract | Crossref Full Text | Google Scholar

48. Jiao, Y, Moon, SJ, and Montell, C. A Drosophila gustatory receptor required for the responses to sucrose, glucose, and maltose identified by mRNA tagging. Proc Natl Acad Sci. (2007) 104:14110–5. doi: 10.1073/pnas.0702421104

PubMed Abstract | Crossref Full Text | Google Scholar

49. Jiao, Y, Moon, SJ, Wang, X, Ren, Q, and Montell, C. Gr64f is required in combination with other gustatory receptors for sugar detection in Drosophila. Curr Biol. (2008) 18:1797–801. doi: 10.1016/j.cub.2008.10.009

PubMed Abstract | Crossref Full Text | Google Scholar

50. Slone, J, Daniels, J, and Amrein, H. Sugar receptors in Drosophila. Curr Biol. (2007) 17:1809–16. doi: 10.1016/j.cub.2007.09.027

PubMed Abstract | Crossref Full Text | Google Scholar

51. Yavuz, A, Jagge, C, Slone, J, and Amrein, H. A genetic tool kit for cellular and behavioral analyses of insect sugar receptors. Fly (Austin). (2014) 8:189–96. doi: 10.1080/19336934.2015.1050569

PubMed Abstract | Crossref Full Text | Google Scholar

52. Dus, M, Min, S, Keene, AC, Lee, GY, and Suh, GS. Taste-independent detection of the caloric content of sugar in Drosophila. Proc Natl Acad Sci USA. (2011) 108:11644–9. doi: 10.1073/pnas.1017096108

PubMed Abstract | Crossref Full Text | Google Scholar

53. Gordesky-Gold, B, Rivers, N, Ahmed, OM, and Breslin, PAS. Drosophila melanogaster prefers compounds perceived sweet by humans. Chem Senses. (2008) 33:301–9. doi: 10.1093/chemse/bjm088

PubMed Abstract | Crossref Full Text | Google Scholar

54. Stafford, JW, Lynd, KM, Jung, AY, and Gordon, MD. Integration of taste and calorie sensing in Drosophila. J Neurosci. (2012) 32:14767–74. doi: 10.1523/JNEUROSCI.1887-12.2012

PubMed Abstract | Crossref Full Text | Google Scholar

55. Lee, Y, Moon, SJ, Montell, C, and Snyder, SH. Multiple gustatory receptors required for the caffeine response in Drosophila. Proc Natl Acad Sci USA. (2009) 106:4495–500. doi: 10.1073/pnas.0811744106

PubMed Abstract | Crossref Full Text | Google Scholar

56. Moon, SJ, Lee, Y, Jiao, Y, and Montell, C. A Drosophila gustatory receptor essential for aversive taste and inhibiting male-to-male courtship. Curr Biol. (2009) 19:1623–7. doi: 10.1016/j.cub.2009.07.061

PubMed Abstract | Crossref Full Text | Google Scholar

57. Shim, J, Lee, Y, Jeong, YT, Kim, Y, Lee, MG, Montell, C, et al. The full repertoire of Drosophila gustatory receptors for detecting an aversive compound. Nat Commun. (2015) 6:8867. doi: 10.1038/ncomms9867

PubMed Abstract | Crossref Full Text | Google Scholar

58. Weiss, LA, Dahanukar, A, Kwon, JY, Banerjee, D, and Carlson, JR. The molecular and cellular basis of bitter taste in Drosophila. Neuron. (2011) 69:258–72. doi: 10.1016/j.neuron.2011.01.001

PubMed Abstract | Crossref Full Text | Google Scholar

59. Leung, NY, Thakur, DP, Gurav, AS, Kim, SH, Di Pizio, A, Niv, MY, et al. Functions of opsins in Drosophila taste. Curr Biol. (2020) 30:e1366:1367–1379.e6. doi: 10.1016/j.cub.2020.01.068

Crossref Full Text | Google Scholar

60. Masuzzo, A, Manière, G, Grosjean, Y, Kurz, L, and Royet, J. Bacteria-derived peptidoglycan triggers a non-canonical NF-κB dependent response in Drosophila gustatory neurons. J Neurosci. (2022) 42:JN-RM-2437-21–823. doi: 10.1523/JNEUROSCI.2437-21.2022

PubMed Abstract | Crossref Full Text | Google Scholar

61. Xiao, S, Baik, LS, Shang, X, and Carlson, JR. Meeting a threat of the Anthropocene: taste avoidance of metal ions by Drosophila. Proc Natl Acad Sci. (2022) 119:e2204238119. doi: 10.1073/pnas.2204238119

PubMed Abstract | Crossref Full Text | Google Scholar

62. Luo, R, Zhang, Y, Jia, Y, Zhang, Y, Li, Z, Zhao, J, et al. Molecular basis and homeostatic regulation of zinc taste. Protein Cell. (2022) 13:462–9. doi: 10.1007/s13238-021-00845-8

PubMed Abstract | Crossref Full Text | Google Scholar

63. Li, X, Sun, Y, Gao, S, Li, Y, Liu, L, and Zhu, Y. Taste coding of heavy metal ion-induced avoidance in Drosophila. iScience. (2023) 26:106607. doi: 10.1016/j.isci.2023.106607

PubMed Abstract | Crossref Full Text | Google Scholar

64. Devineni, AV, Sun, B, Zhukovskaya, A, and Axel, R. Acetic acid activates distinct taste pathways in Drosophila to elicit opposing, state-dependent feeding responses. eLife. (2019) 8:e47677. doi: 10.7554/eLife.47677

PubMed Abstract | Crossref Full Text | Google Scholar

65. Shrestha, B, Aryal, B, and Lee, Y. The taste of vitamin C in Drosophila. EMBO Rep. (2023) 24:e56319. doi: 10.15252/embr.202256319

PubMed Abstract | Crossref Full Text | Google Scholar

66. Stanley, M, Ghosh, B, Weiss, ZF, Christiaanse, J, and Gordon, MD. Mechanisms of lactic acid gustatory attraction in Drosophila. Curr Biol. (2021) 31:3525–3537.e6. doi: 10.1016/j.cub.2021.06.005

PubMed Abstract | Crossref Full Text | Google Scholar

67. Rimal, S, Sang, J, Poudel, S, Thakur, D, Montell, C, and Lee, Y. Mechanism of acetic acid gustatory repulsion in Drosophila. Cell Rep. (2019) 26:1432–1442.e4. doi: 10.1016/j.celrep.2019.01.042

PubMed Abstract | Crossref Full Text | Google Scholar

68. Aryal, B, Dhakal, S, Shrestha, B, and Lee, Y. Molecular and neuronal mechanisms for amino acid taste perception in the Drosophila labellum. Curr Biol. (2022) 32:1376–1386.e4. doi: 10.1016/j.cub.2022.01.060

PubMed Abstract | Crossref Full Text | Google Scholar

69. Guillemin, J, Li, V, Davis, G, Audette, K, Li, J, Jelen, M, et al. Taste cells expressing Ionotropic Receptor 94e reciprocally impact feeding and egg laying in Drosophila. bioRxiv. (2024). doi: 10.1101/2024.01.23.576843

Crossref Full Text | Google Scholar

70. Mi, T, Mack, JO, Koolmees, W, Lyon, Q, Yochimowitz, L, Teng, ZQ, et al. Alkaline taste sensation through the alkaliphile chloride channel in Drosophila. Nat Metab. (2023) 5:466–80. doi: 10.1038/s42255-023-00765-3

PubMed Abstract | Crossref Full Text | Google Scholar

71. Masek, P, and Keene, AC. Drosophila fatty acid taste signals through the PLC pathway in sugar-sensing neurons. PLoS Genet. (2013) 9:e1003710. doi: 10.1371/journal.pgen.1003710

PubMed Abstract | Crossref Full Text | Google Scholar

72. Tauber, JM, Brown, EB, Li, Y, Yurgel, ME, Masek, P, and Keene, AC. A subset of sweet-sensing neurons identified by IR56d are necessary and sufficient for fatty acid taste. PLoS Genet. (2017) 13:e1007059. doi: 10.1371/journal.pgen.1007059

PubMed Abstract | Crossref Full Text | Google Scholar

73. Brown, EB, Shah, KD, Palermo, J, Dey, M, Dahanukar, A, and Keene, AC. Ir56d-dependent fatty acid responses in Drosophila uncover taste discrimination between different classes of fatty acids. eLife. (2021) 10:e67878. doi: 10.7554/eLife.67878

PubMed Abstract | Crossref Full Text | Google Scholar

74. Pradhan, RN, Shrestha, B, and Lee, Y. Molecular basis of Hexanoic acid taste in Drosophila melanogaster. Mol Cells. (2023) 46:451–60. doi: 10.14348/molcells.2023.0035

PubMed Abstract | Crossref Full Text | Google Scholar

75. Sukumaran, SK, Lewandowski, BC, Qin, Y, Kotha, R, Bachmanov, AA, and Margolskee, RF. Whole transcriptome profiling of taste bud cells. Sci Rep. (2017) 7:7595. doi: 10.1038/s41598-017-07746-z

PubMed Abstract | Crossref Full Text | Google Scholar

76. Charroux, B, Daian, F, and Royet, J. Drosophila aversive behavior toward Erwinia carotovora carotovora is mediated by bitter neurons and Leukokinin. iScience. (2020) 23:101152. doi: 10.1016/j.isci.2020.101152

PubMed Abstract | Crossref Full Text | Google Scholar

77. Soldano, A, Alpizar, YA, Boonen, B, Franco, L, López-Requena, A, Liu, G, et al. Gustatory-mediated avoidance of bacterial lipopolysaccharides via TRPA1 activation in Drosophila. eLife. (2016) 5:e13133. doi: 10.7554/eLife.13133

PubMed Abstract | Crossref Full Text | Google Scholar

78. Kouakou, YI, and Lee, RJ. Interkingdom detection of bacterial quorum-sensing molecules by mammalian taste receptors. Microorganisms. (2023) 11:1295. doi: 10.3390/microorganisms11051295

PubMed Abstract | Crossref Full Text | Google Scholar

79. Qin, Y, Palayyan, SR, Zheng, X, Tian, S, Margolskee, RF, and Sukumaran, SK. Type II taste cells participate in mucosal immune surveillance. PLoS Biol. (2023) 21:e3001647. doi: 10.1371/journal.pbio.3001647

PubMed Abstract | Crossref Full Text | Google Scholar

80. Taruno, A, and Gordon, MD. Molecular and cellular mechanisms of salt taste. Annu Rev Physiol. (2023) 85:25–45. doi: 10.1146/annurev-physiol-031522-075853

PubMed Abstract | Crossref Full Text | Google Scholar

81. Lee, Y, Poudel, S, Kim, Y, Thakur, D, and Montell, C. Calcium taste avoidance in Drosophila. Neuron. (2018) 97:e64:67–74.e4. doi: 10.1016/j.neuron.2017.11.038

Crossref Full Text | Google Scholar

82. Mcdowell, SAT, Stanley, M, and Gordon, MD. A molecular mechanism for high salt taste in Drosophila. Curr Biol. (2022) 32:3070–3081.e5. doi: 10.1016/j.cub.2022.06.012

Crossref Full Text | Google Scholar

83. Dweck, HKM, Talross, GJS, Luo, Y, Ebrahim, SAM, and Carlson, JR. Ir56b is an atypical ionotropic receptor that underlies appetitive salt response in Drosophila. Curr Biol. (2022) 32:1776–1787.e4. doi: 10.1016/j.cub.2022.02.063

PubMed Abstract | Crossref Full Text | Google Scholar

84. Zhang, YV, Ni, J, and Montell, C. The molecular basis for attractive salt-taste coding in Drosophila. Science. (2013) 340:1334–8. doi: 10.1126/science.1234133

PubMed Abstract | Crossref Full Text | Google Scholar

85. Lawless, HT, Schlake, S, Smythe, J, Lim, J, Yang, H, Chapman, K, et al. Metallic taste and retronasal smell. Chem Senses. (2004) 29:25–33. doi: 10.1093/chemse/bjh003

Crossref Full Text | Google Scholar

86. Lawless, HT, Stevens, DA, Chapman, KW, and Kurtz, A. Metallic taste from electrical and chemical stimulation. Chem Senses. (2005) 30:185–94. doi: 10.1093/chemse/bji014

PubMed Abstract | Crossref Full Text | Google Scholar

87. Tansaraviput, S, and Nolden, AA. Sucrose, NaCl, and citric acid suppress the metallic sensation of FeSO4. Chem Senses. (2024) 49:bjad052. doi: 10.1093/chemse/bjad052

PubMed Abstract | Crossref Full Text | Google Scholar

88. Clemens, S, and Ma, JF. Toxic heavy metal and metalloid accumulation in crop plants and foods. Annu Rev Plant Biol. (2016) 67:489–512. doi: 10.1146/annurev-arplant-043015-112301

Crossref Full Text | Google Scholar

89. Hou, D, O’connor, D, Igalavithana, AD, Alessi, DS, Luo, J, Tsang, DCW, et al. Metal contamination and bioremediation of agricultural soils for food safety and sustainability. Nat Rev Earth Environ. (2020) 1:366–81. doi: 10.1038/s43017-020-0061-y

Crossref Full Text | Google Scholar

90. Wang, Y, Zajac, AL, Lei, W, Christensen, CM, Margolskee, RF, Bouysset, C, et al. Metal ions activate the human taste receptor TAS2R7. Chem Senses. (2019) 44:339–47. doi: 10.1093/chemse/bjz024

PubMed Abstract | Crossref Full Text | Google Scholar

91. Teng, B, Wilson, CE, Tu, Y-H, Joshi, NR, Kinnamon, SC, and Liman, ER. Cellular and neural responses to sour stimuli require the proton channel Otop1. Curr Biol. (2019) 29:e3645:3647–3656.e5. doi: 10.1016/j.cub.2019.08.077

Crossref Full Text | Google Scholar

92. Zhang, J, Jin, H, Zhang, W, Ding, C, O’keeffe, S, Ye, M, et al. Sour sensing from the tongue to the brain. Cell. (2019) 179:e315:392–402.e15. doi: 10.1016/j.cell.2019.08.031

Crossref Full Text | Google Scholar

93. Ganguly, A, Chandel, A, Turner, H, Wang, S, Liman, ER, and Montell, C. Requirement for an Otopetrin-like protein for acid taste in Drosophila. Proc Natl Acad Sci. (2021) 118:e2110641118. doi: 10.1073/pnas.2110641118

PubMed Abstract | Crossref Full Text | Google Scholar

94. Mi, T, Mack, JO, Lee, CM, and Zhang, YV. Molecular and cellular basis of acid taste sensation in Drosophila. Nat Commun. (2021) 12:3730. doi: 10.1038/s41467-021-23490-5

PubMed Abstract | Crossref Full Text | Google Scholar

95. Da Conceicao Neta, ER, Johanningsmeier, SD, and Mcfeeters, RF. The chemistry and physiology of sour taste—a review. J Food Sci. (2007) 72:R33–8. doi: 10.1111/j.1750-3841.2007.00282.x

Crossref Full Text | Google Scholar

96. Harvey, R. The relation between the total acidity, the concentration of the hydrogen ion, and the taste of acid solutions. J Am Chem Soc. (1920) 42:712–4. doi: 10.1021/ja01449a005

Crossref Full Text | Google Scholar

97. Liman, ER, Zhang, YV, and Montell, C. Peripheral coding of taste. Neuron. (2014) 81:984–1000. doi: 10.1016/j.neuron.2014.02.022

PubMed Abstract | Crossref Full Text | Google Scholar

98. Lyall, V, Alam, RI, Phan, DQ, Ereso, GL, Phan, T-HT, Malik, SA, et al. Decrease in rat taste receptor cell intracellular pH is the proximate stimulus in sour taste transduction. Am J Phys Cell Phys. (2001) 281:C1005–13. doi: 10.1152/ajpcell.2001.281.3.C1005

PubMed Abstract | Crossref Full Text | Google Scholar

99. Charlu, S, Wisotsky, Z, Medina, A, and Dahanukar, A. Acid sensing by sweet and bitter taste neurons in Drosophila melanogaster. Nat Commun. (2013) 4:2042. doi: 10.1038/ncomms3042

PubMed Abstract | Crossref Full Text | Google Scholar

100. Bryant, BP. Mechanisms of somatosensory neuronal sensitivity to alkaline pH. Chem Senses. (2005) 30:i196–7. doi: 10.1093/chemse/bjh182

PubMed Abstract | Crossref Full Text | Google Scholar

101. Kloehn, NW, and Brogden, W. The alkaline taste: a comparison of absolute thresholds for sodium hydroxide on the tip and mid-dorsal surfaces of the tongue. Am J Psychol. (1948) 61:90–3. doi: 10.2307/1417296

PubMed Abstract | Crossref Full Text | Google Scholar

102. Tian, L, Zhang, H, Yang, S, Luo, A, Kamau, PM, Hu, J, et al. Vertebrate OTOP1 is also an alkali-activated channel. Nat Commun. (2023) 14:26. doi: 10.1038/s41467-022-35754-9

PubMed Abstract | Crossref Full Text | Google Scholar

103. Dai, Z, Zheng, W, and Locasale, JW. Amino acid variability, tradeoffs and optimality in human diet. Nat Commun. (2022) 13:6683. doi: 10.1038/s41467-022-34486-0

PubMed Abstract | Crossref Full Text | Google Scholar

104. Ikeda, K. On a new seasoning. J Tokyo Chem Soc. (1909) 30:820–36.

Google Scholar

105. Maga, JA, and Yamaguchi, S. Flavor potentiators. Crit Rev Food Sci Nutr. (1983) 18:231–312. doi: 10.1080/10408398309527364

Crossref Full Text | Google Scholar

106. Toda, Y, Hayakawa, T, Itoigawa, A, Kurihara, Y, Nakagita, T, Hayashi, M, et al. Evolution of the primate glutamate taste sensor from a nucleotide sensor. Curr Biol. (2021) 31:4641–4649.e5. doi: 10.1016/j.cub.2021.08.002

PubMed Abstract | Crossref Full Text | Google Scholar

107. Chaudhari, N, Pereira, E, and Roper, SD. Taste receptors for umami: the case for multiple receptors. Am J Clin Nutr. (2009) 90:738S–42S. doi: 10.3945/ajcn.2009.27462H

PubMed Abstract | Crossref Full Text | Google Scholar

108. Diepeveen, J, Moerdijk-Poortvliet, TC, and Van Der Leij, FR. Molecular insights into human taste perception and umami tastants: a review. J Food Sci. (2022) 87:1449–65. doi: 10.1111/1750-3841.16101

PubMed Abstract | Crossref Full Text | Google Scholar

109. Kohl, S, Behrens, M, Dunkel, A, Hofmann, T, and Meyerhof, W. Amino acids and peptides activate at least five members of the human bitter taste receptor family. J Agric Food Chem. (2013) 61:53–60. doi: 10.1021/jf303146h

PubMed Abstract | Crossref Full Text | Google Scholar

110. Mcgrane, SJ, Gibbs, M, Hernangomez De Alvaro, C, Dunlop, N, Winnig, M, Klebansky, B, et al. Umami taste perception and preferences of the domestic cat (Felis catus), an obligate carnivore. Chem Senses. (2023) 48:bjad026. doi: 10.1093/chemse/bjad026

PubMed Abstract | Crossref Full Text | Google Scholar

111. Nelson, G, Chandrashekar, J, Hoon, MA, Feng, L, Zhao, G, Ryba, NJ, et al. An amino-acid taste receptor. Nature. (2002) 416:199–202. doi: 10.1038/nature726

Crossref Full Text | Google Scholar

112. Pal Choudhuri, S, Delay, RJ, and Delay, ER. L-amino acids elicit diverse response patterns in taste sensory cells: a role for multiple receptors. PLoS One. (2015) 10:e0130088. doi: 10.1371/journal.pone.0130088

PubMed Abstract | Crossref Full Text | Google Scholar

113. San Gabriel, A, Uneyama, H, Yoshie, S, and Torii, K. Cloning and characterization of a novel mGluR1 variant from vallate papillae that functions as a receptor for L-glutamate stimuli. Chem Senses. (2005) 30:i25–6. doi: 10.1093/chemse/bjh095

PubMed Abstract | Crossref Full Text | Google Scholar

114. Zhang, J, Sun-Waterhouse, D, Su, G, and Zhao, M. New insight into umami receptor, umami/umami-enhancing peptides and their derivatives: a review. Trends Food Sci Technol. (2019) 88:429–38. doi: 10.1016/j.tifs.2019.04.008

Crossref Full Text | Google Scholar

115. Toshima, N, and Schleyer, M. Neuronal processing of amino acids in Drosophila: from taste sensing to behavioural regulation. Curr Opin Insect Sci. (2019) 36:39–44. doi: 10.1016/j.cois.2019.07.007

PubMed Abstract | Crossref Full Text | Google Scholar

116. Steck, K, Walker, SJ, Itskov, PM, Baltazar, C, Moreira, J-M, and Ribeiro, C. Internal amino acid state modulates yeast taste neurons to support protein homeostasis in Drosophila. eLife. (2018) 7:e31625. doi: 10.7554/eLife.31625

PubMed Abstract | Crossref Full Text | Google Scholar

117. Kalyanasundar, B, Blonde, GD, Spector, AC, and Travers, SP. Electrophysiological responses to sugars and amino acids in the nucleus of the solitary tract of type 1 taste receptor double-knockout mice. J Neurophysiol. (2020) 123:843–59. doi: 10.1152/jn.00584.2019

PubMed Abstract | Crossref Full Text | Google Scholar

118. Zhang, F, Klebansky, B, Fine, RM, Xu, H, Pronin, A, Liu, H, et al. Molecular mechanism for the umami taste synergism. Proc Natl Acad Sci USA. (2008) 105:20930–4. doi: 10.1073/pnas.0810174106

PubMed Abstract | Crossref Full Text | Google Scholar

119. Koh, TW, He, Z, Gorur-Shandilya, S, Menuz, K, Larter, NK, Stewart, S, et al. The Drosophila IR20a clade of ionotropic receptors are candidate taste and pheromone receptors. Neuron. (2014) 83:850–65. doi: 10.1016/j.neuron.2014.07.012

PubMed Abstract | Crossref Full Text | Google Scholar

120. Kurihara, K. Umami the fifth basic taste: history of studies on receptor mechanisms and role as a food flavor. Biomed Res Int. (2015) 2015:189402:1–10. doi: 10.1155/2015/189402

Crossref Full Text | Google Scholar

121. Rihani, K, Fraichard, S, Chauvel, I, Poirier, N, Delompré, T, Neiers, F, et al. A conserved odorant binding protein is required for essential amino acid detection in Drosophila. Commun Biol. (2019) 2:425. doi: 10.1038/s42003-019-0673-2

PubMed Abstract | Crossref Full Text | Google Scholar

122. Bowen, D, Green, P, Vizenor, N, Vu, C, Kreuter, P, and Rolls, B. Effects of fat content on fat hedonics: cognition or taste? Physiol Behav. (2003) 78:247–53. doi: 10.1016/S0031-9384(02)00973-3

Crossref Full Text | Google Scholar

123. Tsuruta, M, Kawada, T, Fukuwatari, T, and Fushiki, T. The orosensory recognition of long-chain fatty acids in rats. Physiol Behav. (1999) 66:285–8. doi: 10.1016/S0031-9384(98)00299-6

PubMed Abstract | Crossref Full Text | Google Scholar

124. Warwick, ZS, and Synowski, SJ. Effect of food deprivation and maintenance diet composition on fat preference and acceptance in rats. Physiol Behav. (1999) 68:235–9. doi: 10.1016/S0031-9384(99)00192-4

PubMed Abstract | Crossref Full Text | Google Scholar

125. Rolls, ET, Critchley, HD, Browning, AS, Hernadi, I, and Lenard, L. Responses to the sensory properties of fat of neurons in the primate orbitofrontal cortex. J Neurosci. (1999) 19:1532–40. doi: 10.1523/JNEUROSCI.19-04-01532.1999

PubMed Abstract | Crossref Full Text | Google Scholar

126. Pittman, D, Crawley, ME, Corbin, CH, and Smith, KR. Chorda tympani nerve transection impairs the gustatory detection of free fatty acids in male and female rats. Brain Res. (2007) 1151:74–83. doi: 10.1016/j.brainres.2007.03.027

PubMed Abstract | Crossref Full Text | Google Scholar

127. Running, CA, Craig, BA, and Mattes, RD. Oleogustus: the unique taste of fat. Chem Senses. (2015) 40:507–16. doi: 10.1093/chemse/bjv036

PubMed Abstract | Crossref Full Text | Google Scholar

128. Takeda, M, Sawano, S, Imaizumi, M, and Fushiki, T. Preference for corn oil in olfactory-blocked mice in the conditioned place preference test and the two-bottle choice test. Life Sci. (2001) 69:847–54. doi: 10.1016/S0024-3205(01)01180-8

Crossref Full Text | Google Scholar

129. Laugerette, F, Passilly-Degrace, P, Patris, B, Niot, I, Febbraio, M, Montmayeur, J-P, et al. CD36 involvement in orosensory detection of dietary lipids, spontaneous fat preference, and digestive secretions. J Clin Invest. (2005) 115:3177–84. doi: 10.1172/JCI25299

PubMed Abstract | Crossref Full Text | Google Scholar

130. Cartoni, C, Yasumatsu, K, Ohkuri, T, Shigemura, N, Yoshida, R, Godinot, N, et al. Taste preference for fatty acids is mediated by GPR40 and GPR120. J Neurosci. (2010) 30:8376–82. doi: 10.1523/JNEUROSCI.0496-10.2010

PubMed Abstract | Crossref Full Text | Google Scholar

131. Ahn, J-E, Chen, Y, and Amrein, H. Molecular basis of fatty acid taste in Drosophila. eLife. (2017) 6:e30115. doi: 10.7554/eLife.30115

PubMed Abstract | Crossref Full Text | Google Scholar

132. Sánchez-Alcañiz, JA, Silbering, AF, Croset, V, Zappia, G, Sivasubramaniam, AK, Abuin, L, et al. An expression atlas of variant ionotropic glutamate receptors identifies a molecular basis of carbonation sensing. Nat Commun. (2018) 9:4252. doi: 10.1038/s41467-018-06453-1

PubMed Abstract | Crossref Full Text | Google Scholar

133. Kim, H, Kim, H, Kwon, JY, Seo, JT, Shin, DM, and Moon, SJ. Drosophila Gr64e mediates fatty acid sensing via the phospholipase C pathway. PLoS Genet. (2018) 14:e1007229. doi: 10.1371/journal.pgen.1007229

PubMed Abstract | Crossref Full Text | Google Scholar

134. Liu, P, Shah, BP, Croasdell, S, and Gilbertson, TA. Transient receptor potential channel type M5 is essential for fat taste. J Neurosci. (2011) 31:8634–42. doi: 10.1523/JNEUROSCI.6273-10.2011

PubMed Abstract | Crossref Full Text | Google Scholar

Keywords: taste, gustation, chemosensation, Drosophila melanogaster, taste receptor, gustatory receptor, tastant

Citation: Arntsen C, Guillemin J, Audette K and Stanley M (2024) Tastant-receptor interactions: insights from the fruit fly. Front. Nutr. 11:1394697. doi: 10.3389/fnut.2024.1394697

Received: 01 March 2024; Accepted: 02 April 2024;
Published: 11 April 2024.

Edited by:

Gabriella Morini, University of Gastronomic Sciences, Italy

Reviewed by:

Wenli Tian, Chinese Academy of Agricultural Sciences, China

Copyright © 2024 Arntsen, Guillemin, Audette and Stanley. 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: Molly Stanley, molly.stanley@uvm.edu

These authors share first authorship

Download