10
Vol:.(1234567890) Marine Life Science & Technology (2020) 2:6–15 https://doi.org/10.1007/s42995-019-00010-5 1 3 REVIEW Receptor, signal transduction and evolution of sweet, umami and bitter taste Zhongmei Ren 1,2  · Zhenhui Liu 1,2 Received: 11 June 2019 / Accepted: 28 August 2019 / Published online: 25 November 2019 © Ocean University of China 2019 Abstract Like olfaction, the sense of taste allows the detection and discrimination of chemicals in the environment. However, while olfaction is specialized in the detection of volatile chemicals, taste is restricted to the detection of contact-chemicals. Two families of mammalian taste receptors, T1R and T2R, involved in recognition of sweet, umami (the taste of monosodium glutamate) and bitter stimuli have been identified and characterized. Although much progress has been made in studies on the basic mechanisms of taste recognition and signal transduction in mammals, we are still far from a full understanding of different taste qualities. This review presents a current perspective on sweet, bitter and umami taste receptors and their signal transduction mechanism. We also discuss the evolution of taste and taste-related molecules. Keywords Bitter · Evolution · Signal transduction · Sweet · Taste · Umami Introduction Taste enables an animal to evaluate the nutritious content and the overall pleasure of food and thus acts as the last step in food acceptance or rejection behavior. Consequently, taste has evolved as a vital survival mechanism for animals (Call- away 2012). Although a wider range of taste modalities have been proposed, five basic taste qualities are currently recog- nized and widely accepted: sweet, umami, bitter, sour and salty. Each taste modality is mediated by distinct receptors located at the apical pole of the taste receptor cells (TRCs) that are found in taste buds (Barlow 2015; Chandrashekar et al. 2006). Taste receptors for salty and sour stimuli are ion chan- nels (Boughter and Gilbertson 1999; Chandrashekar et al. 2010; Huang et al. 2006; Lindemann 2001; Nelson et al. 2001, 2002). Sweet, bitter and umami receptors belong to the family of G protein-coupled receptors (GPCRs), which possess a seven alpha-helical transmembrane core flanked by an extracellular amino termini and an intracellular carboxy- terminal tail. Here, we review current perspectives on sweet, bitter and umami taste receptors and their signal transduc- tion mechanisms. The evolution of taste is also discussed. Taste buds and taste receptor cells Taste buds, which are peripheral structures responsible for sensing taste compounds in food and drink, are distributed across different papillae of the tongue as well as on the pal- ate epithelium and on the wall of the throat (Miller 1995). There are three kinds of papillae located in the mammalian tongue, i.e., fungiform, foliate and circumvallate (Montma- yeur and Matsunami 2002). The taste bud is an onion-shaped epithelial structure with 60–100 elongate cells. They are classified into at least three morphological types (type I, type II and type III) and as many as five or more functional categories (Feng et al. 2014; Liman et al. 2014). Type I (or glial-like) cells are the most common and are considered to be supporting cells which are wrapped around other cell types and orchestrate re-uptake or degradation of neurotrans- mitters; type II are TRCs for sweet, bitter and umami detec- tion; type III are TRCs that appear to be sour and possibly salty detectors, although the cell type mediating salty taste Edited by Jiamei Li. * Zhenhui Liu [email protected] 1 Department of Marine Biology, Institute of Evolution and Marine Biodiversity, Ocean University of China, Qingdao 266003, China 2 Laboratory for Marine Biology and Biotechnology, Pilot National Laboratory for Marine Science and Technology (Qingdao), Qingdao 266003, China

Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

Vol:.(1234567890)

Marine Life Science & Technology (2020) 2:6–15https://doi.org/10.1007/s42995-019-00010-5

1 3

REVIEW

Receptor, signal transduction and evolution of sweet, umami and bitter taste

Zhongmei Ren1,2 · Zhenhui Liu1,2

Received: 11 June 2019 / Accepted: 28 August 2019 / Published online: 25 November 2019 © Ocean University of China 2019

AbstractLike olfaction, the sense of taste allows the detection and discrimination of chemicals in the environment. However, while olfaction is specialized in the detection of volatile chemicals, taste is restricted to the detection of contact-chemicals. Two families of mammalian taste receptors, T1R and T2R, involved in recognition of sweet, umami (the taste of monosodium glutamate) and bitter stimuli have been identified and characterized. Although much progress has been made in studies on the basic mechanisms of taste recognition and signal transduction in mammals, we are still far from a full understanding of different taste qualities. This review presents a current perspective on sweet, bitter and umami taste receptors and their signal transduction mechanism. We also discuss the evolution of taste and taste-related molecules.

Keywords Bitter · Evolution · Signal transduction · Sweet · Taste · Umami

Introduction

Taste enables an animal to evaluate the nutritious content and the overall pleasure of food and thus acts as the last step in food acceptance or rejection behavior. Consequently, taste has evolved as a vital survival mechanism for animals (Call-away 2012). Although a wider range of taste modalities have been proposed, five basic taste qualities are currently recog-nized and widely accepted: sweet, umami, bitter, sour and salty. Each taste modality is mediated by distinct receptors located at the apical pole of the taste receptor cells (TRCs) that are found in taste buds (Barlow 2015; Chandrashekar et al. 2006).

Taste receptors for salty and sour stimuli are ion chan-nels (Boughter and Gilbertson 1999; Chandrashekar et al. 2010; Huang et al. 2006; Lindemann 2001; Nelson et al.

2001, 2002). Sweet, bitter and umami receptors belong to the family of G protein-coupled receptors (GPCRs), which possess a seven alpha-helical transmembrane core flanked by an extracellular amino termini and an intracellular carboxy-terminal tail. Here, we review current perspectives on sweet, bitter and umami taste receptors and their signal transduc-tion mechanisms. The evolution of taste is also discussed.

Taste buds and taste receptor cells

Taste buds, which are peripheral structures responsible for sensing taste compounds in food and drink, are distributed across different papillae of the tongue as well as on the pal-ate epithelium and on the wall of the throat (Miller 1995). There are three kinds of papillae located in the mammalian tongue, i.e., fungiform, foliate and circumvallate (Montma-yeur and Matsunami 2002). The taste bud is an onion-shaped epithelial structure with 60–100 elongate cells. They are classified into at least three morphological types (type I, type II and type III) and as many as five or more functional categories (Feng et al. 2014; Liman et al. 2014). Type I (or glial-like) cells are the most common and are considered to be supporting cells which are wrapped around other cell types and orchestrate re-uptake or degradation of neurotrans-mitters; type II are TRCs for sweet, bitter and umami detec-tion; type III are TRCs that appear to be sour and possibly salty detectors, although the cell type mediating salty taste

Edited by Jiamei Li.

* Zhenhui Liu [email protected]

1 Department of Marine Biology, Institute of Evolution and Marine Biodiversity, Ocean University of China, Qingdao 266003, China

2 Laboratory for Marine Biology and Biotechnology, Pilot National Laboratory for Marine Science and Technology (Qingdao), Qingdao 266003, China

Page 2: Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

7Marine Life Science & Technology (2020) 2:6–15

1 3

remains ambiguous (Barlow 2015; Barlow and Klein 2015; Chaudhari 2014; Ishimaru et al. 2005; Roper 2007).

The logic of how the taste system encodes information on chemical identity, that is, quality coding, is the subject of active investigation resulting in competing models of taste coding (Ohla et al. 2019; Scott and Giza 2000; Smith and St John 1999; Smith et al. 2000). In the “labeled line” model, the taste cells are selectively tuned to a specific modal-ity, and are innervated by individually tuned nerve fibers. According to this model, the different transmission lines are separate, distinct, and parallel. The sensory afferent neurons are all “specialists” for a given quality. In the “across-fiber pattern” or “combinatorial coding” model, a single taste cell would recognizes a variety of different tastes and respond to multiple taste modalities. This model allows more flexibility in the responses of primary afferent fibers. The molecular and functional studies in mice seem to support the “labeled line” model (Chandrashekar et al. 2006), but more taste receptor cells and genes mediating different taste should be identified to distinguish the two models.

Taste receptors

Two distinct families of taste receptors, T1R and T2R, have been well characterized in mammals. The T1R family belongs to the large class C GPCRs that are easily distin-guished by the presence of long extracellular ligand-binding sites (Kunishima et al. 2000). Receptors in this family con-tain three proteins, T1R1, T1R2, and T1R3, among which T1R2 and T1R3 subunits function together as sweet recep-tors, while T1R1 and T1R3 subunits function together as umami receptors (Adler et al. 2000; Chandrashekar et al. 2000; Hoon et al. 1999; Matsunami et al. 2000; Nelson et al. 2001, 2002; Sainz et al. 2001; Zhao et al. 2003). The dis-covery and identification of T1R as the receptors of sweet or umami stimuli were reviewed by Bachmanov and Beau-champ (2007) and Montmayeur and Matsunami (2002). Briefly, the mouse sac gene was found to encode a sweet taste receptor T1R3 (Bachmanov et al. 1997, 2001; Li et al. 2001, 2002a, b). Studies of functional expression in het-erogeneous cells further showed that T1R3 combined with T1R2 can recognize all kinds of sweet substances includ-ing natural sugars, artificial sweeteners and d-amino acids (Li et al. 2002a, b; Nelson et al. 2001). By contrast, the T1R1/T1R3 heteromeric receptor can detect l-amino acids and monosodium l-glutamate (MSG), which is the taste of umami (Nelson et al. 2002). Subsequently, co-expression of T1R1/T1R3 or T1R2/T1R3 in taste buds and studies on T1R knockout in mice provide powerful evidence for T1R proteins as taste receptors (Damak et al. 2003; Kim et al. 2003; Kiuchi et al. 2006; Zhao et al. 2003).

The T2R family is involved in bitter detection, which may cause aversion and thus a defense mechanism against

ingestion of toxins (Conte et al. 2003; Shi et al. 2003). Die-tary toxins are a major selective force driving the evolution of T2R genes in mammals (Davis et al. 2010; Li and Zhang 2014). The T2R family has a short extracellular amino-ter-minus and is most closely related to the class A (rhodopsin-like) GPCRs, but is regarded as a separate class (Adler et al. 2000; Conte et al. 2003; Matsunami et al. 2000). It contains more than 30 genes in mammals most of which are clus-tered into large linked arrays, suggesting rapid expansion of this family (Adler et al. 2000; Matsunami et al. 2000; Scott 2005). The identification of T2R genes as bitter receptor candidates was largely a result of the release of the human genome (Adler et al. 2000; Matsunami et al. 2000). T2R genes from mouse and rat were identified quickly thereaf-ter (Conte et al. 2002, 2003; Wu et al. 2005). T2R genes are intronless, whereas T1R genes contain multiple introns (Shi and Zhang 2006). Unlike the T1R receptor family, there is no evidence that the T2R receptor functions as a dimer, but the possibility of T2R as a hybrid cannot be ruled out. In addition, some T2Rs detect only one bitter compound, while others react to more than 50 natural and synthetic bitter chemicals (Behrens et al. 2014; Lossow et al. 2016; Trivedi 2012a).

The receptors for bitter, sweet and umami are not limited to the tongue—they are also found in the intestines, stom-ach, pancreas, respiratory tract and even sperm (Roper 2013; Spector et al. 1993; Trivedi 2012b; Xu et al. 2013). Studying the function of these taste receptors, which are distributed throughout the body, could help clinicians tackle diseases ranging from eating disorders to diabetes (Trivedi 2012b). In addition to their sensory function, TRCs release a variety of endocrine-active substances, such as glucagon-like peptides, ghrelin, leptin and neuropeptide Y, which are involved in regulating food intake (Calvo and Egan 2015).

Do more types of taste receptors exist?

It has been known that T1R2/T1R3 recognizes all sugar substances, TIR1/T1R3 is the umami receptor, and T2Rs is the bitter receptor. Are there any more taste receptors to detect sweet/umami/bitter substances? For example, is there a T1R1/T1R2 heterodimer or homodimer? If so, what are their functions? Kim et al. (2003) have found that, either alone or in all possible combinations, members of the T1R family was present in both circumvallate and fungiform papillae, and that T1R1 and T2R were coexpressed, in mice. Liao and Schultz (2003) showed that all three T1Rs were present selectively in taste receptor cells in human fungiform papillae. The existence of multiple sweet receptors was sup-ported by electrophysiological and cross-adaptation studies (Danilova and Hellekant 2003; Froloff et al. 1998; Ninomiya et al. 1993; Schiffman et al. 1981). Furthermore, Damak et al. (2003) suggested that mice lacking the T1R3 receptor

Page 3: Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

8 Marine Life Science & Technology (2020) 2:6–15

1 3

could not taste artificial sweeteners but still had an affin-ity for sugars, especially glucose, and could also perceive umami (Damak et al. 2003; Trivedi 2012a). This indicates that there are two different sweet mechanisms. However, no other members of the T1Rs family have been found by extensive searches of the close-to-complete human genome (Liao and Schultz 2003). Potential taste receptors may have no obvious sequence homology with T1Rs or T2Rs.

There may exist other bitter receptor candidates besides the T2R family. Some studies have shown that bitter com-pounds such as caffeine and theophylline interact directly with phosphodiesterase (PDE) in taste tissue (Rosenzweig et al. 1999). Thus, it is speculated that the bitterness mecha-nism of caffeine is produced by inhibiting PDE (Rosenzweig et al. 1999). In addition, G protein-dependent and -independ-ent pathways for denatonium signal transduction both exist in mammalian taste cells (Sawano et al. 2005). Therefore, more receptors that do not belong to GPCRs may be found in G protein-independent pathways.

Tongue map and brain map

Does the tongue have a “map” that responds to the five basic tastes of sweet, bitter, umami, salty and sour? For example, does circumvallate on the back of the tongue respond only to bitter, while foliate on the sides of the tongue recognize only sour? Recent molecular and functional data have revealed that all five basic modalities are present in all the taste papillae of the tongue (Adler et al. 2000; Hoon et al. 1999; Huang et al. 2006; Nelson et al. 2001, 2002). So, contrary to the popular belief in the early 1900s, there is no specific taste “map” on the tongue. However, each taste bud on the tongue responds to only one taste (Callaway 2012; Chen et al. 2011).

How does the brain transform detection into perception? One of the leading theories is that the gustatory cortex was ‘broadly tuned’ with each neuron responding particularly well to one taste but still responding to others. Chen et al. (2011), however, found that sweet, bitter, salty and umami each was represented in its own separate cortical field, revealing the existence of a gustotopic map in the brain. Nevertheless, there may be some sensory crossover, with some cells in the bitter hotspot, for example, also responding to other tastes (Trivedi 2012b).

Taste signal transduction

Neuroscientists who study taste are just beginning to under-stand how and why the interaction of a few molecules on the tongue can trigger taste behaviors. Cells devoted to the detection of sweet, umami, and bitter stimuli share common signal transduction components (Meyerhof 2005). Generally, the first step is the ligand’s interaction with a specific GPCR

on the membrane. The G protein-mediated signaling path-way is then activated, initiating two second-messenger sign-aling cascades: cyclic adenosine monophosphate (cAMP) and phosphoinositide signaling, which are two parallel streams of intracellular events. Finally, depolarization of taste cells occurs, which activates nerve fibers (Gilbertson and Boughter 2003). Thus, taste signals are conveyed to the brain stem where central taste processing begins, eliciting adaptive responses (for details see review of Gilbertson and Boughter 2003; review of Lee and Owyang 2017; review of Lindemann 2001; review of Roper 2007).

Here, we focused on the early events of taste signal trans-duction including the ligands, receptors, G proteins, scaffold proteins and their interactions.

Ligands

It is apparent that different species enjoy different food com-pounds, and preferences differ even within the same spe-cies, for example, in humans. This leads us to consider the specificities of taste ligands. There is some research on this, although not very systematic. For example, Nelson et al. (2002) have shown that human T1R2/T1R3 receptors can recognize aspartame, cyclamate and various sweet proteins, but mouse T1R2/T1R3 cannot.

Ligands recognized as taste receptors have been largely determined by in vitro heterologous expression and in vivo experiments. Human T1R2/T1R3 are known to recognize diverse natural sugars, such as sucrose and glucose, as well as synthetic sweeteners, such as saccharin and acesulfame-K (Li et al. 2002a, b; Roper 2007). Human T1R1/T1R3 functions as a broadly turned l-amino-acid sensor respond-ing to most of the 20 standard amino acids, but not to their d-enantiomers (Nelson et al. 2002). When the taste stimulus is l-glutamate or monosodium glutamate (MSG), inosine 5-monophosphate (IMP) or guanosine 5-ribonucleotide (GMP) can enhance the response, which is the hallmark of umami taste (Li et al. 2002a, b; Roper 2007). More than 30 T2R bitter taste receptors have been discovered in mam-mals (Adler et al. 2000; Matsunami et al. 2000). The ago-nists for various T2Rs have also been identified and all of which were established as bitter tastants. Several receptors have been found to sense numerous bitter substances (Bufe et al. 2002). This is in line with the finding that the number of human bitter compounds is much larger than the num-ber of human T2R genes. However, it was also found that different T2Rs selectively recognize different bitter com-pounds. For example, the human hT2R4 and mouse mT2R8 (a probable hT2R4 orthologue) is activated by denatonium (Chandrashekar et al. 2000); mouse mT2R5 (Chandrashekar et al. 2000) and rat rT2R9s (mT2R5 homologues; Bufe et al. 2002) respond to cycloheximide; human receptors T2R10 and T2R16 respond to strychnine and salicin, respectively

Page 4: Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

9Marine Life Science & Technology (2020) 2:6–15

1 3

(Bufe et al. 2002). Moreover, Mueller et al. (2005) demon-strated that mice lacking the receptor of mT2R5 lose the ability to recognize cycloheximide but can still detect other bitter compounds.

The binding of ligands and receptors

Given that T1R or T2R respond to a large wide range of taste compounds, one concept emerging is that there are multi-ple binding pockets with different selectivities for different ligands on the receptors (Roper 2007). Nevertheless, the question remains how the thousands of structurally diverse taste substances can be recognized by different combination pockets. It is more likely that receptors are broadly tuned or that other mechanisms exist.

The large extracellular domain of T1Rs is the candidate site of critical binding pockets for the detection and dis-crimination of ligands. Walters (2002) analyzed the large extracellular domain of the sweet receptor T1R3 using a homology-modeled method based on the crystal structure of the metabotropic glutamate receptor (mGluR1). It showed that the regions corresponding to mGluR1 ligand-binding site in the model would respond to polyhydroxy compounds such as monosaccharides and disaccharides. Furthermore, using computer modeling, Temussi (2006) demonstrated that the large extracellular domain of T1R dimer forms a hinged ligand-binding pocket (also known as Venus flytrap, VFT). Experimental studies have confirmed that the large extracel-lular domains of T1R2 and T1R3 bind sugars, sweet amino acids and other sweet compounds (Morini et al. 2005). In addition, when the large extracellular N-terminal domain

of T1Rs is expressed as soluble protein, the ligand-binding function remains (Nie et al. 2006).

Further binding sites have been identified using chimeric constructs or site-directed mutations. By heterologous expression of the chimeras of mouse and human T1R1 and T1R3, Jiang et al. (2004) showed that a small area in the cysteine-rich region near the base of the N-terminus of T1R3 was required for receptor activity toward brazzein. Xu et al. (2004) demonstrated that the N-terminal VFT domain of T1R2 was required for the recognition of sweeteners such as aspartame and neotame, while the C-terminal trans-membrane domain of T1R3 was required for the recogni-tion of sweetener cyclamate and sweet inhibitor lactisole. Jiang et al. (2005) further proposed that the transmembrane domain of human T1R3 was likely to play a critical role in the interconversion of the sweet receptor from the ground state to the active state. They also identified several resi-dues within the transmembrane domain of human T1R3 that account in large part for the species-specific response to cyclamate. Figure 1 shows the multiple binding sites between the GPCRs and ligands.

Unlike T1Rs, bitter receptor T2Rs has a shorter extracel-lular N-terminal, so its role in ligand recognition may not be as important as that of T1Rs. By domain swapping between hT2R43, hT2R44 and hT2R47, it was suggested that the extracellular loop 1 (EC1) of T2R43 was critical for its bind-ing with IMNB (N-isopropyl-2-methyl-5-nitrobenzenesul-fonamide) and 6-nitrosaccharin (Pronin et al. 2004). In addi-tion, the extracellular loop 2 (EC2) was involved in binding to 6-nitrosaccharin, while the third extracellular loop (EC3) was not (Pronin et al. 2004). The structure–function relation-ships of bitter receptors hTAS2R46, -R43, and -R31 with

Fig. 1 Schematic drawing of the multiple binding sites between sweet and umami taste receptors and ligands (based on the data of Xu et al. 2004; Roper 2007). The N-terminal extracellular domain of T1R2 is required for the binding of aspartame, neotame, saccharin and sucrose. The transmembrane domains of T1R3 are required for the bind-ing of cyclamate, lactisole, sac-charin and artificial sweeteners. The N-termini of T1R3 near the first transmembrane region is required for certain sweet-tast-ing proteins such as brazzein. l-Glutamate and IMP bind to the T1R1 subunit, respectively. In addition, the transmembrane domain of T1R1 or TIR2 is responsible for coupling to G proteins

Page 5: Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

10 Marine Life Science & Technology (2020) 2:6–15

1 3

their agonists were also revealed by calcium imaging of chi-meric and mutant receptors (Brockhoff et al. 2007, 2010). It was found that the carboxyl-terminal regions, extracellular loops and transmembrane domains were all crucial for ago-nist selectivity (Brockhoff et al. 2007, 2010).

G proteins

Heterotrimeric guanine-nucleotide-binding proteins (G pro-teins), consisting of Gα and Gβγ subunits, function as signal transducers of the seven transmembrane helix of GPCRs. The specific interaction between GPCRs and heterotrimeric G proteins is important for maintaining the specificity and fidelity of signal relay (Gudermann et al. 1996; Neves et al. 2002). However, the native taste signaling pathways involv-ing G protein are only poor understood. McLaughlin et al. (1992) identified a novel G protein α-subunit, α-gustducin, from taste tissue. The expression of α-gustducin messenger RNA was found in taste buds of all taste papillae including circumvallate, foliate and fungiform, but not in non-sensory portions of the tongue or other tissues, indicating a role of α-gustducin in taste transduction (McLaughlin et al. 1992). The role of α-gustducin in sweet, bitter and umami transduc-tion was subsequently verified by comparing the behavioral and electrophysiological responses in α-gustducin knock-out and wild-type mice (Caicedo et al. 2003; Danilova et al. 2006; He et al. 2004; Ruiz et al. 2003; Wong et al. 1996), or by transgenic expression of α-gustducin in α-gustducin-null mice (He et al. 2002). It has been demonstrated that other G protein subunits, i.e., Gβ3 and Gγ13, are coexpressed with α -gustducin in taste receptor cells, indicating the possible role of Gβ3 and Gγ13 in taste signal transduction (Huang et al. 1999, 2003; Max et al. 2001).

How do the receptors interact with G proteins? Xu et al. (2004) demonstrated that T1R2, rather than T1R3, was criti-cal for Gα15 coupling, requiring the transmembrane domain of T1R2 (Fig. 1). For T2Rs, the cytoplasmic loops and adjacent transmembrane segments have the greatest con-servation, which is the predicted site for G protein binding. Thirty-seven C-terminal amino acids of gustducin and Gαi2 are indispensable for the T2R activity (Ueda et al. 2003). In addition, there are species differences in the map of receptor and G protein interactions. The G protein-coupling efficiency differs even between humans and rats (Xu et al. 2004).

The role of scaffold proteins

The role of scaffold proteins in taste signal transduction is poorly understood. Meyer et al. (2002) found that scaffold proteins such as ZO-1, ZO-2 and ZO-3 could interact with signaling protein Gα12, indicating a possible role in signal transduction. ZO-1/ZO-2 has also been shown to interact

with tight junction protein claudins. Claudins were found to be present specifically in mouse taste papillae and human fungiform papillae (Michlig et al. 2007). Furthermore, ZO-1 can interact with Gγ13, a G protein gamma 13 subunit that plays a key role in taste signal transduction. ZO-1 and Gγ13 are co-located in taste bud cells (Liu et al. 2012). Therefore, scaffold proteins may be involved in taste signal transduction (Fig. 2), although precise mechanism by which this occurs is not yet clear.

Taste evolution

Taste perception varies enormously across different spe-cies of animals. In birds, chickens did not seem to respond to sweet, while other nectar feeders showed strong prefer-ences for sweet substances (Ganchrow et al. 1990; Shi and Zhang 2006); In mammals, different trophic groups differ in their selectivity and sensitivity to bitter or sweet com-pounds (Glendinning 1994; Xu et al. 2004). For example, a cat cannot taste sweet substances. Even between rats and mice, there were differences in electrophysiological proper-ties, expression of markers and innervation in taste buds (Richter et al. 2004; Zaidi and Whitehead 2006). Ma et al. (2007) further suggested that there was a significant differ-ence in the percentage of taste cells expressing signaling molecules between rats and mice. In fact, to adapt to differ-ent environments, taste organs have evolved to accommodate a distinct function in different species or lineages (Scott et al. 2001). For example, in vertebrates, taste is largely restricted to the tongue and palate, while in insects, gustatory neurons are more broadly distributed along the body, not only in the proboscis and pharynx, but also in the wings, legs, and female genitalia (Scott et al. 2001).

Fig. 2 The role of scaffold protein ZO-1/ZO-2 complex in taste sig-nal transduction. ZO-1 interacts with Gγ13 and Gα12 through PDZ domain 1 and SH3 domain, respectively; it associates with clau-dins, ZO-2 and JAM (junctional adhesion molecules) through PDZ domains 1, 2, and 3, respectively (Ebnet et  al. 2004; Fanning et  al. 1998; Itoh et al. 2001); it associates with occludin through the GUK domain (Fanning et  al. 1998) and with F-actin through an actin-binding region in the C-terminal half of the molecule (Fanning et al. 2002). The binding interactions among other tight junction proteins are not shown, and whether all of the interactions occur simultane-ously remain to be determined

Page 6: Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

11Marine Life Science & Technology (2020) 2:6–15

1 3

Taste perception is mediated by taste receptor cells, in which taste receptors trigger the first step of taste signal transduction. Taste receptors specify attraction or aversion to food, so changes in these receptors may be directly related to adaptive changes in diet (Liman 2006). Taste receptors have now been identified in a large number of species includ-ing humans, other primates, mice, rats, fish and flies (Conte et al. 2003; Fischer et al. 2005; Go et al. 2005; Scott et al. 2001; Shi et al. 2003; Wang et al. 2004), enabling us to gain a better understanding of their evolutionary history and their associated functions.

Shi and Zhang (2006) performed comparative evolu-tionary analyses of T1R and T2R genes from nine verte-brates. Three T1R genes were found in humans, mice, rats, dogs and opossums, but the number of T1R genes varies in some non-mammalian vertebrates. For example, fugu and pufferfish have five and six T1R genes, respectively (Shi and Zhang 2006). No T1R2 genes were found in chickens, which may account for their lack of sweet taste (Ganchrow et al. 1990; Shi and Zhang 2006). Amphibians (for exam-ple, western clawed frog) may have lost all T1R genes, but this requires confirmation (Shi and Zhang 2006). Cats have a mutation that disables one of the two genes that build a working sweet receptor, so cats cannot taste sweet substances. It is likely that the sweet-receptor gene became inactive in their evolutionary ancestor (Callaway 2012; Li et al. 2005). Jiang et al. (2012) showed that among 12 non-feline species in the order Carnivora, six species also had unique mutations in the sweet taste receptor gene (Jiang et al. 2012), suggesting that the ability to taste sweetness had been lost repeatedly over the course of evolution. In addition, dolphins lack the ability to build working umami and bitter taste receptors and pandas, a largely vegetarian member of Carnivora, lack umami receptors (Callaway 2012). There are four to six T2R genes in fish and three in chickens, but 21–42 in mammals and 49 in frogs (Shi and Zhang 2006), indicating a species- or lineage-specific gene duplication event. In mammals, the evolutionary relation-ships of T2R gene families have been described in detail (Conte et al. 2003; Fischer et al. 2005; Go et al. 2005; Lei et al. 2015; Sandau et al. 2015; Wang et al. 2004). Five T2R protein groups existed prior to the divergence of the primate and rodent lineages (Conte et al. 2003; Shi et al. 2003), and the species- or lineage-specific gene duplica-tion and pseudogenization were shown to play an impor-tant role in altering of the T2R gene repertoire in indi-vidual primate species (Go et al. 2005; Zhao et al. 2010). When analyzing the rate of non-synonymous substitution versus synonymous substitution, positive selection of T1R and T2R genes was detected in vertebrates (Fischer et al. 2005; Shi and Zhang 2006; Shi et al. 2003; Wang et al. 2004). These two gene families, however, exhibited drasti-cally different modes of evolution, with the former being

conserved and latter being radical in the changes of gene number and gene sequences (Shi and Zhang 2006). Thus, functional divergence and specialization of taste receptors in vertebrates generally occur via adaptive evolution (Shi and Zhang 2006).

It has long been known that mice and humans perceive various sweet substances through a single T1R2/T1R3 heterodimer (Li et al. 2002a, b; Zhao et al. 2003). Fish, however, appear to have different taste reception rules (Hashiguchi et al. 2007; Oike et al. 2007). Several T1R2 genes have been characterized in fish, all of which act as receptors to amino acids but not to sugars (Hashigu-chi et al. 2007; Oike et al. 2007). Moreover, the T1R2 sequence similarity between fish and mammalian species is not high (Go 2006; Ishimaru et al. 2005; Pfister and Rodrigue 2005; Shi and Zhang 2006). This indicates that T1R2 is a changeable gene and several T1R2 genes may have existed before the separation of teleost fish and tet-rapods (Oike et al. 2007). In the process of fish evolution, T1R2 gene replication events may have occurred giving T1Rs fish-specific functions such as a high sensitivity to amino acids. By contrast, mammalian T1R2/T1R3 recep-tors may have acquired sugar-responding ability (Hashi-guchi et al. 2007; Oike et al. 2007).

To date, T1R has been found only in fish, amphibians, birds and mammals, but not in invertebrates (Oike et al. 2007). No genes coding for bitter, sweet and umami taste receptors were found in the amphioxus Branchiostoma flor-idae (subphylum Cephalochordata) or the tunicate Ciona intestinalis (subphylum Urochordata), although olfactory receptors were identified in the former (Churcher and Tay-lor, 2009; Nordström et al. 2008). Interestingly, a large and diverse family of 68 candidate gustatory receptor (GR) genes was found in Drosophila (Clyne et al. 2000; Dunipace et al. 2001; Robertson et al. 2003; Scott et al. 2001). The carboxyl terminus of all GRs contains a signature motif which is also present in some members of the Drosophila odorant recep-tor (DOR) family, suggesting that they may share a com-mon evolutionary origin. Scott et al. (2001) speculated that the GR gene family was likely to encode both olfactory and gustatory receptors in Drosophila, but not in vertebrates. Toshima and Tanimura (2012) proposed the presence of amino acid receptors and a monitoring system that regulates the feeding responses to amino acids in D. melanogaster. In addition, by comparing the molecular evolution of the GR genes along D. sechellia and D. simulans lineages, a rapid evolution of taste receptor genes during host specialization in D. sechellia was found. Furthermore, D. sechellia is los-ing GR genes nearly ten times faster than its generalist sib-ling D. simulans (McBride 2007). It is therefore likely that GR genes are subject to novel evolutionary pressures when insects enter new niches during host shifts or host specializa-tion events (McBride 2007).

Page 7: Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

12 Marine Life Science & Technology (2020) 2:6–15

1 3

Future directions

In recent years, with the discovery of T1R and T2R recep-tors, much progress has been achieved in the research of sweet, umami and bitter taste. However, many fundamen-tal issues remain to be resolved. For example, exactly how many taste receptors exist for different taste qualities? How do receptors recognize their specific ligands in taste bud cells and what are the pathways of taste signal trans-duction? How have taste receptors evolved? The answers to these basic questions will help us understand how taste works.

Acknowledgements This work was supported by the grants of National Natural Science Foundation of China (31572259) and the Fundamen-tal Research Funds for the Central Universities of China (201822020; 201841013).

Author contributions ZR wrote the manuscript; ZL provided the out-line for this manuscript and improved the text.

Compliance with ethical standards

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

Animal and human rights statement This paper followed the ethical guidelines for animal and human rights established by Ocean Univer-sity of China (Permit Number, SD2007695).

References

Adler E, Hoon MA, Mueller KL, Chandrashekar J, Ryba NJ, Zuker CS (2000) A novel family of mammalian taste receptors. Cell 100:693–702

Bachmanov AA, Beauchamp GK (2007) Taste receptor genes. Annu Rev Nutr 27:389–414

Bachmanov AA, Reed DR, Ninomiya Y, Inoue M, Tordoff MG, Price RA, Beauchamp GK (1997) Sucrose consumption in mice: major influence of two genetic loci affecting peripheral sensory responses. Mamm Genome 8:545–548

Bachmanov AA, Li X, Reed DR, Ohmen JD, Li S, Chen Z, Tordoff MG, De Jong PJ, Wu C, West DB, Chatterjee A, Ross DA, Beau-champ GK (2001) Positional cloning of the mouse saccharin preference (Sac) locus. Chem Senses 26:925–933

Barlow LA (2015) Progress and renewal in gustation: new insights into taste bud development. Development 142:3620e9

Barlow LA, Klein OD (2015) Developing and regenerating a sense of taste. Curr Top Dev Biol 111:401–419

Behrens M, Korsching SI, Meyerhof W (2014) Tuning properties of avian and frog bitter taste receptors dynamically fit gene reper-toire sizes. Mol Biol Evol 31:3216–3227

Boughter JD, Gilbertson TA (1999) From channels to behavior: an integrative model of NaCl taste. Neuron 22:213–215

Brockhoff A, Behrens M, Massarotti A, Appendino G, Meyerhof W (2007) Broad tuning of the human bitter taste receptor hTAS2R46 to various sesquiterpene lactones, clerodane and

labdane diterpenoids, strychnine, and denatonium. J Agric Food Chem 55:6236–6243

Brockhoff A, Behrens M, Niv MY, Meyerhof W (2010) Structural requirements of bitter taste receptor activation. Proc Natl Acad Sci USA 107:11110–11115

Bufe B, Hofmann T, Krautwurst D, Raguse JD, Meyerhof W (2002) The human TAS2R16 receptor mediates bitter taste in response to beta-glucopyranosides. Nat Genet 32:397–401

Caicedo A, Pereira E, Margolskee RF, Roper SD (2003) Role of the G-protein subunit alpha-gustducin in taste cell responses to bitter stimuli. J Neurosci 23:9947–9952

Callaway E (2012) The lost appetites. Nature 486:S16–S17Calvo SS, Egan JM (2015) The endocrinology of taste receptors. Nat

Rev Endocrinol 11:213–227Chandrashekar J, Mueller KL, Hoon MA, Adler E, Feng L, Guo W,

Zuker CS, Ryba NJ (2000) T2Rs function as bitter taste recep-tors. Cell 100:703–711

Chandrashekar J, Hoon MA, Ryba NJ, Zuker CS (2006) The recep-tors and cells for mammalian taste. Nature 444:288–294

Chandrashekar J, Kuhn C, Oka Y, Yarmolinsky DA, Hummler E, Ryba NJ, Zuker CS (2010) The cells and peripheral representa-tion of sodium taste in mice. Nature 464:297–301

Chaudhari N (2014) Synaptic communication and signal process-ing among sensory cells in taste buds. J Physiol 592:3387e92

Chen X, Gabitto M, Peng Y, Ryba NJ, Zuker CS (2011) A gusto-topic map of taste qualities in the mammalian brain. Science 333:1262–1266

Churcher AM, Taylor JS (2009) Amphioxus (Branchiostoma flori-dae) has orthologs of vertebrate odorant receptors. BMC Evol Biol 9:242

Clyne PJ, Warr CG, Carlson JR (2000) Candidate taste receptors in Drosophila. Science 287:1830–1834

Conte C, Ebeling M, Marcuz A, Nef P, Andres-barquin PJ (2002) Identification and characterization of human taste receptor genes belonging to the TAS2R family. Cytogenet Genome Res 98:45–53

Conte C, Ebeling M, Marcuz A, Nef P, Andres-barquin PJ (2003) Evolutionary relationships of the Tas2r receptor gene families in mouse and human. Physiol Genomics 14:73–82

Damak S, Rong M, Yasumatsu K, Kokrashvili Z, Varadarajan V, Zou S, Jiang P, Ninomiya Y, Margolskee RF (2003) Detection of sweet and umami taste in the absence of taste receptor T1r3. Science 301:850–853

Danilova V, Hellekant G (2003) Comparison of the responses of the chorda tympani and glossopharyngeal nerves to taste stimuli in C57BL/6 J mice. BMC Neurosci 4:5

Danilova V, Damak S, Margolskee RF, Hellekant G (2006) Taste responses to sweet stimuli in alpha-gustducin knockout and wild-type mice. Chem Senses 31:573–580

Davis JK, Lowman JJ, Thomas PJ (2010) Evolution of a bitter taste receptor gene cluster in a new world sparrow. Genome Biol Evol 3:358–370

Dunipace L, Meister S, Mcnealy C, Amrein H (2001) Spatially restricted expression of candidate taste receptors in the Dros-ophila gustatory system. Curr Biol 11:822–835

Ebnet K, Suzuki A, Ohno S, Vestweber D (2004) Junctional adhe-sion molecules (JAMs): more molecules with dual functions? J Cell Sci 117:19–29

Fanning AS, Jameson BJ, Jesaitis LA, Anderson JM (1998) The tight junction protein ZO-1 establishes a link between the transmem-brane protein occludin and the actin cytoskeleton. J Biol Chem 273:29745–29753

Fanning AS, Ma TY, Anderson JM (2002) Isolation and functional characterization of the actin binding region in the tight junction protein ZO-1. Faseb J 16:1835–1837

Page 8: Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

13Marine Life Science & Technology (2020) 2:6–15

1 3

Feng P, Huang L, Wang H (2014) Taste bud homeostasis in health, disease and aging. Chem Senses 39:3–16

Fischer A, Gilad Y, Man O, Pääbo S (2005) Evolution of bitter taste receptors in humans and apes. Mol Biol Evol 22:432–436

Froloff N, Lloret E, Martinez JM, Faurion A (1998) Cross-adaptation and molecular modeling study of receptor mechanisms common to four taste stimuli in humans. Chem Senses 23:197–206

Ganchrow JR, Steiner JE, Bartana A (1990) Behavioral reactions to gustatory stimuli in young chicks (Gallus gallus domesticus). Dev Psychobiol 23:103–117

Gilbertson TA, Boughter JD (2003) Taste transduction: appetizing times in gustation. NeuroReport 14:905–911

Glendinning JI (1994) Is the bitter rejection response always adaptive? Physiol Behav 56:1217–1227

Go Y (2006) Lineage-specific expansions and contractions of the bit-ter taste receptor gene repertoire in vertebrates. Mol Biol Evol 23:964–972

Go Y, Satta Y, Takenaka O, Takahata N (2005) Lineage-specific loss of function of bitter taste receptor genes in humans and nonhuman primates. Genetics 170:313–326

Gudermann T, Kalkbrenner F, Schultz G (1996) Diversity and selec-tivity of receptor-G protein interaction. Annu Rev Pharmacol 36:429–459

Hashiguchi Y, Furuta Y, Kawahara R, Nishida M (2007) Diversifica-tion and adaptive evolution of putative sweet taste receptors in threespine stickleback. Gene 396:170–179

He W, Danilova V, Zou S, Hellekant G, Max M, Margolskee RF, Damak S (2002) Partial rescue of taste responses of α-gustducin null mice by transgenic expression of α-gustducin. Chem Senses 27:719–727

He W, Yasumatsu K, Varadarajan V, Yamada A, Lem J, Ninomiya Y, Margolskee RF, Damak S (2004) Umami taste responses are mediated by alpha-transducin and alpha-gustducin. J Neurosci 24:7674–7680

Hoon MA, Adler E, Lindemeier J, Battey JF, Ryba NJ, Zuker CS (1999) Putative mammalian taste receptors: a class of taste-specific GPCRs with distinct topographic selectivity. Cell 96:541–551

Huang L, Shanker YG, Dubauskaite J, Zheng JZ, Yan W, Rosenzweig S, Spielman AI, Max M, Margolskee RF (1999) Ggamma13 colocalizes with gustducin in taste receptor cells and mediates IP3 responses to bitter denatonium. Nat Neurosci 2:1055–1062

Huang L, Max M, Margolskee RF, Su H, Masland RH, Euler T (2003) G protein subunit G gamma 13 is coexpressed with G alpha o, G beta 3, and G beta 4 in retinal ON bipolar cells. J Comp Neurol 455:1–10

Huang AL, Chen X, Hoon MA, Chandrashekar J, Guo W, Tränkner D, Ryba NJ, Zuker CS (2006) The cells and logic for mammalian sour taste detection. Nature 442:934–938

Ishimaru Y, Okada S, Naito H, Nagai T, Yasuoka A, Matsumoto I, Abe K (2005) Two families of candidate taste receptors in fishes. Mech Develop 122:1310–1321

Itoh M, Sasaki H, Furuse M, Ozaki H, Kita T, Tsukita S (2001) Junc-tional adhesion molecule (JAM) binds to PAR-3: a possible mechanism for the recruitment of PAR-3 to tight junctions. J Cell Biol 154:491–497

Jiang P, Ji Q, Liu Z, Snyder LA, Benard LM, Margolskee RF, Max M (2004) The cysteine-rich region of T1R3 determines responses to intensely sweet proteins. J Biol Chem 279:45068–45075

Jiang P, Cui M, Zhao B, Snyder LA, Benard LM, Osman R, Max M, Margolskee RF (2005) Identification of the cyclamate interaction site within the transmembrane domain of the human sweet taste receptor subunit T1R3. J Biol Chem 280:34296–34305

Jiang P, Josue J, Li X, Glaser D, Li W, Brand JG, Margolskee RF, Reed DR, Beauchamp GK (2012) Major taste loss in carnivorous mammals. Proc Natl Acad Sci USA 109:4956–4961

Kim MR, Kusakabe Y, Miura H, Shindo Y, Ninomiya Y, Hino A (2003) Regional expression patterns of taste receptors and gustducin in the mouse tongue. Biochem Biophys Res Commun 312:500–506

Kiuchi S, Yamada T, Kiyokawa N, Saito T, Fujimoto J, Yasue H (2006) Genomic structure of swine taste receptor family 1 member 3, TAS1R3, and its expression in tissues. Cytogenet Genome Res 115:51–61

Kunishima N, Shimada Y, Tsuji Y, Sato T, Yamamoto M, Kumasaka T, Nakanishi S, Jingami H, Morikawa K (2000) Structural basis of glutamate recognition by a dimeric metabotropic glutamate receptor. Nature 407:971–977

Lee AA, Owyang C (2017) Sugars, sweet taste receptors, and brain responses. Nutrients 9:653

Lei W, Ravoninjohary A, Li X, Margolskee RF, Reed DR, Beauchamp GK, Jiang P (2015) Functional analyses of bitter taste receptors in domestic cats (Felis catus). PLoS ONE 10:e0139670

Li D, Zhang J (2014) Diet shapes the evolution of the vertebrate bitter taste receptor gene repertoire. Mol Biol Evol 31:303–309

Li X, Inoue M, Reed DR, Huque T, Puchalsk RB, Tordoff MG, Ninomiya Y, Beauchamp GK, Bachmanov AA (2001) High-resolution genetic mapping of the saccharin preference locus (Sac) and the putative sweet taste receptor (T1R1) gene (Gpr70) to mouse distal Chromosome 4. Mamm Genome 12:13–16

Li X, Bachmanov AA, Li S, Chen Z, Tordoff MG, Beauchamp GK, De Jong PJ, Wu C, Chen L, West DB, Ross DA, Ohmen JD, Reed DR (2002a) Genetic, physical, and comparative map of the subtelomeric region of mouse Chromosome 4. Mamm Genome 13:5–19

Li X, Staszewski L, Xu H, Durick K, Zoller M, Adler E (2002b) Human receptors for sweet and umami taste. Proc Natl Acad Sci USA 99:4692–4696

Li X, Li W, Wang H, Cao J, Maehashi K, Huang L, Bachmanov AA, Reed DR, Legrand-defretin V, Beauchamp GK, Brand JG (2005) Pseudogenization of a sweet-receptor gene accounts for cats’ indifference toward sugar. PLoS Genet 1:27–35

Liao J, Schultz PG (2003) Three sweet receptor genes are clustered in human chromosome 1. Mamm Genome 14:291–301

Liman ER (2006) Use it or lose it: molecular evolution of sensory signaling in primates. Pflug Arch Eur J Phy 453:125–131

Liman ER, Zhang YV, Montell C (2014) Peripheral coding of taste. Neuron 81:984–1000

Lindemann B (2001) Receptors and transduction in taste. Nature 413:219–225

Liu Z, Fenech C, Cadiou H, Grall S, Tili E, Laugerette F, Wiencis A, Grosmaitre X, Montmayeur JP (2012) Identification of new binding partners of the chemosensory signaling protein Gγ13 expressed in taste and olfactory sensory cells. Front Cell Neu-rosci 6:26

Lossow K, Hubner S, Roudnitzky N, Slack JP, Pollastro F, Behren M, Meyerhof W (2016) Comprehensive analysis of mouse bit-ter taste receptors reveals different molecular receptive ranges for orthologous receptors in mice and humans. J Biol Chem 291:15358–15377

Ma H, Yang R, Thomas SM, Kinnamon JC (2007) Qualitative and quantitative differences between taste buds of the rat and mouse. BMC Neurosci 8:5

Matsunami H, Montmayeur JP, Buck LB (2000) A family of candidate taste receptors in human and mouse. Nature 404:601–604

Max M, Shanker YG, Huang L, Rong M, Liu Z, Campagne F, Wein-stein H, Damak S, Margolskee RF (2001) Tas1r3, encoding a new candidate taste receptor, is allelic to the sweet responsive-ness locus Sac. Nat Genet 28:58–63

McBride CS (2007) Rapid evolution of smell and taste receptor genes during host specialization in Drosophila sechellia. Proc Natl Acad Sci USA 104:4996–5001

Page 9: Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

14 Marine Life Science & Technology (2020) 2:6–15

1 3

McLaughlin SK, Mckinnon PJ, Margolskee RF (1992) Gustducin is a taste-cell-specific G protein closely related to the transducins. Nature 357:563–569

Meyer TN, Schwesinger C, Denker BM (2002) Zonula occludens-1 is a scaffolding protein for signaling molecules. Galpha(12) directly binds to the Src homology 3 domain and regulates paracellular permeability in epithelial cells. J Biol Chem 277:24855–24858

Meyerhof W (2005) Elucidation of mammalian bitter taste. Rev Physiol Bioch P 154:37–72

Michlig S, Damak S, Le Coutre J (2007) Claudin-based permeability barriers in taste buds. J Comp Neurol 502:1003–1011

Miller IJ (1995) Anatomy of the peripheral taste system. In: Doty RL (ed) Handbook of olfaction and gustation. Marcel Dekker, New York, pp 521–547

Montmayeur JP, Matsunami H (2002) Receptors for bitter and sweet taste. Curr Opin Neurobiol 12:366–371

Morini G, Bassoli A, Temussi PA (2005) From small sweeteners to sweet proteins: anatomy of the binding sites of the human T1R2_T1R3 receptor. J Med Chem 48:5520–5529

Mueller KL, Hoon MA, Erlenbach I, Chandrashekar J, Zuker CS, Ryba NJ (2005) The receptors and coding logic for bitter taste. Nature 434:225–229

Nelson G, Hoon MA, Chandrashekar J, Zhang Y, Ryba NJ, Zuker CS (2001) Mammalian sweet taste receptors. Cell 106:381–390

Nelson G, Chandrashekar J, Hoon MA, Feng L, Zhao G, Ryba NJ, Zuker CS (2002) An amino-acid taste receptor. Nature 416:199–202

Neves SR, Ram PT, Iyengar R (2002) G protein pathways. Science 296:1636–1639

Nie Y, Hobbs JR, Vigues S, Olson WJ, Conn GL, Munger SD (2006) Expression and purification of functional ligand-binding domains of T1R3 taste receptors. Chem Senses 31:505–513

Ninomiya Y, Kajiura H, Mochizuki K (1993) Differential taste responses of mouse chorda tympani and glossopharyngeal nerves to sugars and amino acids. Neurosci Lett 163:197–200

Nordström KJ, Fredriksson R, Schiöth HB (2008) The amphioxus (Branchiostoma floridae) genome contains a highly diversified set of G protein-coupled receptors. BMC Evol Biol 8:9

Ohla K, Yoshida R, Roper SD, Di Lorenzo PM, Victor JD, Boughter JD, Fletcher M, Katz DB, Chaudhari N (2019) Recognizing taste: coding patterns along the neural axis in mammals. Chem Senses 44:237–247

Oike H, Nagai T, Furuyama A, Okad S, Aihara Y, Ishimaru Y, Maru T, Matsumoto I, Misaka T, Abe K (2007) Characterization of ligands for fish taste receptors. J Neurosci 27:5584–5592

Pfister P, Rodrigue I (2005) Olfactory expression of a single and highly variable V1r pheromone receptor-like gene in fish spe-cies. Proc Natl Acad Sci USA 102:5489–5494

Pronin AN, Tang H, Connor J, Keung W (2004) Identification of ligands for two human bitter T2R receptors. Chem Senses 29:583–593

Richter TA, Dvoryanchikov GA, Roper SD, Chaudhari N (2004) Acid-sensing ion channel-2 is not necessary for sour taste in mice. J Neurosci 24:4088–4091

Robertson HM, Warr CG, Carlson JR (2003) Molecular evolution of the insect chemoreceptor gene superfamily in Drosophila melanogaster. Proc Natl Acad Sci USA 100:14537–14542

Roper SD (2007) Signal transduction and information processing in mammalian taste buds. Pflug Arch Eur J Phy 454:759–776

Roper SD (2013) Taste buds as peripheral chemosensory processors. Semin Cell Dev Biol 24:71–79

Rosenzweig S, Yan W, Dasso M, Spielman AI (1999) Possible novel mechanism for bitter taste mediated through cGMP. J Neuro-physiol 81:1661–1665

Ruiz CJ, Wray K, Delay E, Margolskee RF, Kinnamon SC (2003) Behavioral evidence for a role of alpha-gustducin in glutamate taste. Chem Senses 28:573–579

Sainz E, Korley JN, Battey JF, Sullivan SL (2001) Identification of a novel member of the T1R family of putative taste receptors. J Neurochem 77:896–903

Sandau MM, Goodman JR, Thomas A, Rucker JB, Rawson NE (2015) A functional comparison of the domestic cat bitter receptors Tas2r38 and Tas2r43 with their human orthologs. BMC Neurosci 16:33

Sawano S, Seto E, Mori T, Hayashi Y (2005) G protein-dependent and -independent pathways in denatonium signal transduction. Biosci Biotech Bioch 69:1643–1651

Schiffman SS, Cahn H, Lindley MG (1981) Multiple receptor sites mediate sweetness: evidence from cross adaptation. Pharmacol Biochem Behav 15:377–388

Scott K (2005) Taste recognition: food for thought. Neuron 48:455–464

Scott TR, Giza BK (2000) Issues of gustatory neural coding: where they stand today. Physiol Behav 69:65–76

Scott K, Brady R, Cravchik A, Morozov P, Rzhetsky A, Zuker C, Axel R (2001) A chemosensory gene family encoding can-didate gustatory and olfactory receptors in Drosophila. Cell 104:661–673

Shi P, Zhang J (2006) Contrasting modes of evolution between ver-tebrate sweet/umami receptor genes and bitter receptor genes. Mol Biol Evol 23:292–300

Shi P, Zhang J, Yang H, Zhang YP (2003) Adaptive diversification of bitter taste receptor genes in mammalian evolution. Mol Biol Evol 20:805–814

Smith DV, St John SJ (1999) Neural coding of gustatory information. Curr Opin Neurobiol 9:427–435

Smith DV, John SJ, Boughter JD (2000) Neuronal cell types and taste quality coding. Physiol Behav 69:77–85

Spector AC, Travers SP, Norgren R (1993) Taste receptors on the anterior tongue and nasoincisor ducts of rats contribute syner-gistically to behavioral responses to sucrose. Behav Neurosci 107:694–702

Temussi P (2006) The history of sweet taste: not exactly a piece of cake. J Mol Recognit 19:188–199

Toshima N, Tanimura T (2012) Taste preference for amino acids is dependent on internal nutritional state in Drosophila mela-nogaster. J Exp Biol 215:2827–2832

Trivedi B (2012a) The finer points of taste. Nature 486:S2–S3Trivedi B (2012b) Hardwired for taste. Nature 486:S7–S9Ueda T, Ugawa S, Yamamura H, Imaizumi Y, Shimada S (2003)

Functional interaction between T2R taste receptors and G-pro-tein alpha subunits expressed in taste receptor cells. J Neurosci 23:7376–7380

Walters DE (2002) Homology-based model of the extracellu-lar domain of the taste receptor T1R3. Pure Appl Chem 74:1117–1123

Wang X, Thomas SD, Zhang J (2004) Relaxation of selective con-straint and loss of function in the evolution of human bitter taste receptor genes. Hum Mol Genet 13:2671–2678

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

Wu SV, Chen MC, Rozengurt E (2005) Genomic organization, expression, and function of bitter taste receptors (T2R) in mouse and rat. Physiol Genomics 22:139–149

Xu H, Staszewski L, Tang H, Adler E, Zoller M, Li X (2004) Differ-ent functional roles of T1R subunits in the heteromeric taste receptors. Proc Natl Acad Sci USA 101:14258–14263

Page 10: Reptor, ansduc v eet,...10 Marine Life Science & Technology (2020) 2:6–15 1 3 theiragonistswerealsorevealedbycalciumimagingofchi-mericandmutantreceptors(Brockhoetal.2007, …

15Marine Life Science & Technology (2020) 2:6–15

1 3

Xu J, Cao J, Iguchi N, Riethmacher D, Huang L (2013) Functional characterization of bitter-taste receptors expressed in mam-malian testis. Mol Hum Reprod 19:17–28

Zaidi FN, Whitehead MC (2006) Discrete innervation of murine taste buds by peripheral taste neurons. J Neurosci 26:8243–8253

Zhao GQ, Zhang Y, Hoon MA, Chandrashekar J, Erlenbach I, Ryba NJ, Zuker CS (2003) The receptors for mammalian sweet and umami taste. Cell 115:255–266

Zhao H, Yang JR, Xu H, Zhang J (2010) Pseudogenization of the umami taste receptor gene Tas1r1 in the giant panda coincided with its dietary switch to bamboo. Mol Biol Evol 27:2669–2673