You are on page 1of 10

R: Concise Reviews

JFS R: Concise Reviews and Hypotheses in Food Science

in Food Science
Sweet Taste in Man: A Review
B. MEYERS AND M.S. BREWER

ABSTRACT: A greater understanding of the molecular mechanisms of sweet taste has profound significance for
the food industry as well as for consumers. Understanding the mechanism by which sweet taste is elicited by
saccharides, peptides, and proteins will assist science and industry in their search for sweet substances with fewer
negative health effects. The original AH-B theories have been supplanted by detailed structural models. Recent iden-
tification of the human sweet receptor as a dimeric G-protein coupled receptor comprising T1R2 and T1R3 subunits
has greatly increased the understanding of the mechanisms involved in sweet molecule binding and sweet taste
transduction. This review discusses early theories of the sweet receptor, recent research of sweetener chemorecep-
tion of nonprotein and protein ligands, homology modeling, the transduction pathway, the possibility of the sweet
receptor functioning allosterically, as well as the implications of allelic variation.
Keywords: chemoreception, protein sweeteners, saccharides, sweetness receptor, sweetness transduction

Introduction the sweet taste produced by various ligands. The AH-B theory for

F ive taste modalities have been identified in humans: sweet,


sour, bitter, salty, and umami (Temussi 2006a). Over the course
of human evolution, two of these modalities have proven most
sweetener/sweetener binding site interaction was one of the most
widely accepted models. Initially proposed by Shallenberger and
Acree (1967), this model proposed that a sweet-tasting compound
significant to the survival of the species: sweetness provides a must contain a hydrogen bond donor (AH) as well as a hydrogen
means to seek out an energy source in the form of carbohy- bond acceptor (B). At a distance of 2.5 to 4 Å, the AH-B unit on the
drates, and bitterness results in an aversion to alkaloids and sweet molecule (glycophore), can react with a complimentary AH-
other potentially deadly toxins found in nature (Cui and others B unit on the receptor, forming a pair of hydrogen bonds (Figure 1).
2006). Shallenberger and Acree (1967) proposed that sweet tasting amino
Consuming sweet foods can have a significant effect on body acids, which greatly differ in chemical structure from saccharide
mass, overall health, and healthcare-related expenses. Over the sweeteners, exhibit a spatial barrier, of given distance, perpendic-
past 30 y, the prevalence of overweight and obese individuals in the ular to the receptor that allows for attachment of specific side-
United States has increased significantly (Flegal and Troiano 2000). chain conformations while excluding others. These side-chain–
Per capita healthcare expenditures are 81% higher for morbidly receptor interactions were the suggested potentiators of the sweet
obese adults than for adults of normal weight (Arterburn and others taste of these compounds (Shallenberger 1996; Eggers and others
2005). This increase in obesity parallels an increase in energy intake 2000).
(Salbe and others 2004). A greater understanding of the sweet taste Expanding on the work of Shallenberger and Acree (1967), Kier
chemoreception mechanism, the way in which sweet molecules in- (1972) proposed the addition of a 3rd component, “x” (later referred
teract with the sweet receptor, and sweet taste transduction, the to as “γ ”), to the glycophore model which modulates sweet potency
downstream signaling pathway triggered by sweetener-stimulation of the bound ligand by means of hydrophobic interactions. This
of the receptor, could provide food scientists with the knowledge potency-modulating effect occurs due to the hydrophobic group’s
required to develop new sweeteners, lower-calorie food products, effect on the electric potential on the AH-B subunit. Although this
and potentially, a healthier population. model expanded on the previous theory by accounting for differ-
Early theories of sweetener–receptor interactions (prior to 2001) ences in sweetener potencies, it did not specifically require the
were purely speculative because they were based on the structure of hydrophobic component on the glycophore to trigger the transduc-
sweeteners themselves. The discovery of the sweet taste receptor, a tion pathway (Eggers and others 2000).
dimeric G-protein coupled receptor (GPCR; Max and others 2001; Nofre and Tinti (1996) proposed a much more complex model
Nelson and others 2001; Li and others 2002; Zhao and others 2003), comprising 8 functional categories, organized into high affinity
has led to dramatic gains in recent years. This article reviews the and secondary sites, which contribute to sweet taste. Labeled
current and past theories of sweet taste in man focusing on sweet B-, AH-, XH-, G1-, G2-, G3-, G4-, and D, these 8 recognition sites
taste chemoreception and transduction. interact with 8 sweetener interaction sites of the same names, re-
spectively (Figure 2). This theory suggested that profound con-
The AH-B theories and the sweetener “glycophore” formational changes occur within the receptor and its binding
Prior to the discovery of the human sweet taste receptor, sev- site(s) due to hydrogen bonding. The number of binding sites
eral research groups attempted to explain the mechanism behind involved dictates the potency of the sweetener (Nofre and Tinti
1996). Goodman and others (2002) theorized that the aromatic
interactions with the D zone of the sweet receptor are responsi-
MS 20080071 Submitted 1/28/2008, Accepted 4/25/2008. Author Meyers is
with The NutraSweet Co., Chicago, IL 60654, U.S.A. Author Brewer is with ble for the intense sweetness of peptide-based ligands (neotame,
the Dept. of Food Science and Human Nutrition, Univ. of Illinois, 202 ABL, aspartame).
1302 West Pennsylvania Ave., Urbana, IL 61801, U.S.A. Direct inquiries to The aforementioned models of the sweet receptor–binding site
author Brewer (E-mail: msbrewer@uiuc.edu).
interaction are inferentially based on the structure of the binding


C 2008 Institute of Food Technologists
R
Vol. 73, Nr. 6, 2008—JOURNAL OF FOOD SCIENCE R81
doi: 10.1111/j.1750-3841.2008.00832.x
Further reproduction without permission is prohibited
R: Concise Reviews Sweet taste review . . .
in Food Science

ligands, not on the structure of the sweet receptor itself (Cui and GPCR. T1R3 was identified as the likely Sac candidate as it is the
others 2006). Each of the aforementioned models assumes a sin- only such receptor in the identified syntenic region. In addition, it
gle sweetener orthosteric (binding) site; however, more than 1 such shares a specific polymorphism that differentiates all sweet-taster
site has been demonstrated in the human sweet taste receptor (see strains of mice from nontaster strains.
below; Dubois 2004). The commonality shared by all of the indi- The T1R2 and T1R3 receptors are often co-expressed (as a
rect molecular models is the presence of the AH-B glycophore and, complex) in the mammalian palate. Based on changes in Ca2+ con-
in fact, these hydrogen donor/acceptor groups have been shown centration, Nelson and others (2001) reported responses to sweet
to exist within 2 active sites of the sweet taste receptor (Temussi compounds, including sucrose, fructose, saccharin, acesulfame-K,
2006a). dulcin, GA-1, and G1-2 by this complex; cells lacking either the
T1R2 or T1R3 subunit proved nonresponsive.
Discovery of the sweet taste receptor Subsequently, recognition of several classes of sweeteners by
Paving the way for elucidation of human sweet taste receptor the human T1R2/T1R3 receptor was demonstrated. Because the
was the characterization of the mouse sweet taste receptor. Hoon G-protein that couples to the human T1Rs in vivo was unknown,
and others (1999) described a class of taste-specific mammalian human T1R receptors were transfected into a human embryonic
GPCRs. Subsequently, the Sac locus for mice, the chromosomal kidney (HEK293) derived cell line, which expresses a known G-
region that determines the sensing of sweet taste, was identified protein. These receptors (T1R2/T1R3) responded to sugars (su-
and genetically mapped (Li and others 2001). The specific receptor crose, fructose, galactose, glucose, lactose, maltose), amino acids
within the mouse Sac locus responsible for sweet taste was isolated (glycine, D-trypotophan), sweet proteins (monellin, thaumatin),
from circumvallate papillae of the mouse tongue. The distribution and high potency sweeteners (acesulfame potassium, aspartame,
of these T1R3 receptors on the tongue differed from that of the cyclamate, dulcin, neotame, saccharin, sucralose; Li and others
other T1R class receptors. Additionally, genetic mapping showed 2002).
that T1R3 resides on the mouse Sac locus (Kitagawa and others Zhao and others (2003) demonstrated that sweet taste (and
2001). Taken together, these findings made a strong case that T1R3 concurrently umami taste) depended on the presence of T1R-
was a mouse sweet taste receptor candidate. receptors. Genetically modified animals not expressing T1R sub-
Sainz and others (2001) validated that not only was T1R3’s units lost both taste modalities. T1R2 and T1R3 knockout mice ex-
expression taste receptor cell-specific (TRC) but also alleic vari- hibited severely impaired sweet taste perception. However, high
ation of T1R3 occurred between taster and nontaster strains of saccharide concentrations were still able to elicit a detectable re-
mice. Montmayeur and others (2001) demonstrated that very sponse in both strains, indicating that T1R2 and T1R3 receptors
diverse T1R3 expression occurred in the mouse cirumvallate, fo- may be able to function independently, albeit less effectively, as
liate, and fungiform taste papillae. More important, however, was sweet taste receptors for a selective class of sweet compounds.
the observation that most cells expressing T1R3 also express T1R2, Taken together, the studies by Max and others (2001), Nelson and
indicating that the sweet receptor may, in fact, function as a others (2001), Li and others (2002), and Zhao and others (2003)
heterodimer. demonstrate that the T1R2/T1R3 complex is likely responsible for
However, the rodent sweet taste receptor does not respond to the reception of sweet ligands in humans.
several sweeteners that elicit responses in humans. Rat sweet taste
receptors respond to glucose, fructose, maltose, lactose, galactose, Structure of the sweet taste receptor
dulcin, saccharin, acesulfame potassium, sucralose, D-tryptophan, The sweet receptor is defined as a class C GPCR that exists as
and glycine, but not to aspartame, cyclamate, monellin, neotame, a heterodimer of the T1R2 and T1R3 subunits (Figure 3; Cui and
or thaumatin (Li and others 2002). To allow for a greater under- others 2006). Many receptors in the human nervous system are
standing of the human sweet receptor, it was imperative to identify coupled to different classes of G-proteins. More than 50% of drugs
the human Sac locus. on the market regulate the activity of GPCRs (Fredholm and oth-
Searching the human genome for a region syntenic to the mouse ers 2007). In mammals, there are 7 subfamilies of GPCRs: A, B,
Sac locus, Max and others (2001) identified a previously unknown LNB-7TM, C, Frizzled/Smoothened, taste 2, and Vomeronasal type
1. Classes A, B, C, and Vomeronasal type 1 receptors all involve

G2 G5
G4

G1 G3

XH AH

Figure 1 --- Schematic representation of the AH-B gly- Figure 2 --- The Nofre-Tinti model for the sweet receptor
cophore (adapted from Shallenberger and Acree 1967). (adapted from Nofre and Tinti 1996).

R82 JOURNAL OF FOOD SCIENCE—Vol. 73, Nr. 6, 2008


R: Concise Reviews
Sweet taste review . . .

in Food Science
activation of the coupled G-protein to trigger a signal cascade. Modeling the sweet taste receptor
GPCRs are coupled to a guanosine triphosphate (GTP) binding pro- The complete crystal structure of the human T1R2–T1R3 com-
tein that acts as an intermediary between enzymes and ion chan- plex has yet to be solved. Therefore, currently, the only way to ex-
nels. Two classes of these G-proteins have been identified: het- amine ligand binding by this receptor is by examining homology
erotrimeric G-proteins and cytoplasmic G-proteins. Heterotrimeric of models. The crystal structure of the NTD for the class C GPCR
G-proteins, composed of α, β, and γ subunits, act with GPCRs in glutamate receptor, mGluR1, has been identified (Kunishima and
the transduction pathway. To date, 28 α subunits, 5 β subunits, and others 2000). There are 5 crystalline structures of this homodimer
12 γ subunits have been identified. However, the β and γ subunits known. Although mGluR1 exhibits only a 20% similarity to T1R re-
are regarded as 1 combined functional subunit. Based on their de- ceptors (Temussi 2006b), mGluR1 homology model constructs of
gree of similarity in the α subunits, the heterotrimeric G-proteins T1Rs appear quite frequently in the literature. While better models
fall into 1 of 4 families (Naim and others 2006) of the receptor are still required, much has been gained using the
Class C GPCRs, including the sweet receptor, are composed mGluR1 homolgy model. Nie and others (2006) recently reported
of 3 domains: a large extracellular amino terminal or “N” termi- expression and purification of functional ligand-binding domains
nal domain, often referenced in the literature as the “Venus fly- of T1R3 taste receptors. The NTD of mouse T1R3 was expressed
trap” domain; ATD, NTD, or VFTD (NTD), a cysteine-rich domain and purified; subsequent spectroscopic analysis confirmed that the
(CRD); and a 7 helices transmembrane domain (7TM or TMD). The protein was folded and capable of binding ligands. These findings
CRD is composed of approximately 70 amino acids and acts as a will, no doubt, have profound effects on research into T1R3-ligand
bridge between the NTD and 7TMD (Figure 3). The stability of this binding. Future studies will likely use this purification method to
dimeric complex is the result of disulphide bonds between cysteine support or disprove current theories of T1R3 NTD ligand binding
residues in the large extracellular domain; these cysteine residues based on homology models.
available for disulphide bonding do not exist in much more com- Until recently, no crystal structure had been solved for any class
mon class A GPCRs (Fredholm and others 2007). C GPCR CRD. Yu and others (2004) previously proposed a homology
The human sweet taste receptor is coupled to a G αi protein model based on the structure of the tumor necrosis factor receptor
(Ozeck and others 2004). Stimulation of bitter, sweet, and umami (TNFR; Cui and others 2006). The amino acid sequence of TNFR
receptors in HEK293 cells correlates to protein kinase phospho- and CRDs of T1R3 share only 15% identity, yet they share 7 of 9 cys-
rylation and cyclic adenosine monophosphate (cAMP) concentra- teine residues, indicating that the structures may have similar fold-
tions. Pertussis toxin, isolated from Bortaldella pertussis, is a known ing patterns. However, most recently, the entire crystal structure
inhibitor of G αi proteins. When HEK293 cells are treated with per- of the extracellular region of the group II mGluR subtype 3 recep-
tussis toxin, the signals are inhibited. Because protein kinase phos- tor, which includes the CSD, was reported (Muto and others 2007).
phorylation and cAMP are 2nd messengers of many GPCRs and Future studies will also likely use this crystal structure to support
because treatment of HEK293 cells with pertusis toxin inhibit these or disprove current theories of CRD interactions of bound ligands
signals, Ozeck and others (2004) have concluded that sweet, bitter, based on homology models.
and umami receptor cells must be coupled to the G αi protein used The complete crystal structure of bovine rhodopsin, which falls
in their study. into subfamily A GPCR receptors, has been used extensively as
Interestingly, the T1R2, but not the T1R3, subunit plays an im- a template for homology modeling of different classes of GPCRs.
portant role coupling to the G α15 protein in the human sweet taste Cherezov and others (2007) recently reported the crystal struc-
receptor. When the C terminus domain of the rat T1R2 subunit was ture of an engineered β 2 -adrenergic GCPR (β 2 AR), which also falls
substituted in place of the human C-terminus, lack of a response into subfamily A. These researchers note that a number of mod-
to Ace-K and sucrose proved that there was no coupling of the G α15 els of β 2 AR, based on the crystal structure of bovine rhodopsin,
protein (Xu and others 2004). However, replacing the same domain are reported in the literature (Bissantz and others 2003; Furse and
of the T1R3 subunit did not result in the same response. Therefore, Lybrand 2003; Freddolino and others 2004; Gouldson and others
although both T1R2 and T1R3 subunits are necessary for binding of 2004; Zhang and others 2006). However, it is further noted that all
different sweet ligands, only the T1R2 subunit is required for both of these models share a stronger similarity with bovine rhodopsin
ligand recognition and G-protein coupling. than with that of β 2 AR. The authors, therefore, caution against the

Figure 3 --- Schematic representation


of the hT1R2-hT1R3 sweet taste
receptor, including the ATD, CRD,
and 7TMD regions of the
heterodimer complex (adapted from
Cui and others 2006).

Vol. 73, Nr. 6, 2008—JOURNAL OF FOOD SCIENCE R83


R: Concise Reviews Sweet taste review . . .
in Food Science

use of a single structural template for homology modeling and as- identified type A binding sites were able to host only 4 compounds:
sert that the use of a 2nd model will result in more accurate predic- saccharin, alitame, aspartame, and 6-Cl-D-tryptophan. However,
tions. Cui and others (2006) have suggested that a homology model both B type models had binding affinities for a much greater
of bovine rhodopsin can act as a template for the 7TMD of the hu- number of small molecular weight compounds. B chain models
man sweet taste receptor. Perhaps a homology model will be pro- also had binding affinities for high molecular weight proteins in a
posed using both bovine rhodopsin and β 2 AR as a template for the large wedge-like cavity. These results seem to validate the Wedge
human 7TMD and will therefore be a more accurate approximation Model first proposed by Temussi (2002) wherein high molecular
of its structure. weight proteins can be bound in a large wedge-like cavity.
Even though a complete crystal structure of the T1R2/T1R3 re- In contrast to the work of Walters (2002), Xu and others (2004)
ceptor is not currently available, homology models for each of the have demonstrated that the N-terminal NTD of the T1R2 subunit,
3 regions of the receptor have been proposed based on the crystal not that of the T1R3, is responsible for binding the sweet peptide,
structures of the NTD for the protein mGluR1, the TNFR, and the aspartame, and neotame (Table 1). When a human/rat chimera
7TMD of bovine rhodopsin (Cui and others 2006). Furthermore, the composed of the human T1R2 subunit N-terminal domain and
crystal structures of the CRD of the group II mGluR subtype 3 recep- rat T1R3 subunit was used, a response to both ligands occurred.
tor and the GCPR β 2 AR have been solved; these discoveries should The normal rat T1R2–T1R3 complex does not respond to either
result in even more accurate models of these 2 regions of the human sweet compound; therefore, the N-terminal extracellular do-
TRC in the future. By combining GPCR homology models, a tem- main of the human T1R2 must be responsible for binding these
plate for the human sweet taste receptor can be approximated (Cui ligands.
and others 2006). Additionally, the NTD of mouse T1R3 has been In silico modeling based on the mGluR1-NTD crystal structure
isolated and purified. predicted a binding mechanism for aspartame confirming the
hypothesis proposed by Xu and others (2004). Three primary
Nonprotein ligand binding sites interactions were identified between aspartame and hT1R2 NTD:
of the T1R2/T1R3 receptor salt bridges (Asp278 and Asp307), hydrogen bonds (Ser303, Arg383,
Homology modeling based on the protein mGluR1 has been and Val384), and hydrophobic interactions (Tyr215, Tyr103, and
used to evaluate binding mechanisms for the extracellular domain Pro277). In addition, 2 water molecules are required to bridge the
of the T1R3 sweet taste receptor. The protein mGluR1 is a dimeric aspartame molecule in the binding pocket (Cui and others 2006).
receptor existing in open and closed conformations (Walters 2002). Binding sites for the saccharide sweeteners sucrose and glucose
In the presence of a bound glutamate ligand, the 2 monomers, or were proposed for the NTD of both T1R2 and T1R3 subunits. Bind-
ligand binding lobes, assume a closed formation (Figure 4). Using ing affinities were determined using purified mouse NTD proteins
in silico modeling of the mGluR1 dimeric receptor, a mechanism for both receptor subunits. Although both subunits exhibited an
for binding 3 ultra-high potency sweet compounds, neotame, affinity for these saccharides, T1R3 bound sucrose more readily
superaspartame, and SC-45647, was proposed. Polar side chains whereas T1R2 bound glucose more readily (Nie and others 2005).
in the cleft between the monomers readily interact with hydroxyl Recently, a candidate-binding site for cyclamate was identi-
groups on these sweeteners as well as on mono- and disaccharides fied on the human sweet receptor (Jiang and others 2005c). The
(Walters 2002). human receptor (hT1R2-hT1R3) responds to cyclamate, whereas
The extracellular domain of the T1R2 and T1R3 subunits has the mouse receptor (mT1R2-mT1R3) does not. Therefore, mixed
been modeled using both A and B chains of the mGluR1 template species pairings of the 2 sweet receptor subunits were used to
(inactive open–open form and active closed–open form, respec- determine the specific cyclamate-sensitive region. The hT1R3
tively). Using in silico free energy ligand binding calculations, subunit was required to elicit a sweet taste receptor response. Di-
Morini and others (2005) identified at least 4 possible binding rected mutagenesis, in addition to molecular modeling, identified
sites for all classes of sweet compounds. A binding site for small 6 residues in the 7TMD exclusively responsible for binding cycla-
molecular weight compounds was identified in each of the 2 NTDs. mate. Furthermore, when the cyclamate-docking site was modeled,
A Wedge site for high molecular weight proteins as well as a binding overlap was observed for the sweet taste inhibitor lactisole with the
site for allosteric modulators was identified in the 7TMD. The binding site (see subsequently). Perhaps when lactisole is bound
to the 7TMD hT1R3 subunit, cyclamate binding is not possible due
to competition for the binding site. Using a homology model of the
hT1R3 TMD based on the bovine rhodospin crystal structure, Cui
and others (2006) demonstrated a binding pocket located among
TM-3,5,6.

Table 1 --- Summary of binding sites of nonprotein sweet-


eners commonly used in foods.
Sweetener Binding site(s)
Sucrose T1R2 NTD, T1R3 NTD
Glucose T1R2 NTD, T1R3 NTD
Sucralose T1R2 NTD, T1R3 NTD
Saccharin T1R2 NTD, T1R3 NTD
Acesulfame-K T1R2 NTD, T1R3 NTD
Cyclamate T1R3 7TMD
Open Closed Aspartame T1R2 NTD
Neotame T1R2 NTD
Figure 4 --- The mGluR1 receptor protein in both “open: Preferential binding sites (agonist) are shown in bold. Adapted from Bishay
and “closed” conformations (adapted from Walters 2002). (2006).

R84 JOURNAL OF FOOD SCIENCE—Vol. 73, Nr. 6, 2008


R: Concise Reviews
Sweet taste review . . .

in Food Science
Sweet inhibitors such as lactisole are also likely to bind to the Efforts to elucidate similarities in structure between the 3 sweet
human sweet taste receptor. Galindo-Cuspinera and Breslin (2006) proteins brazzein, monellin, and thaumatin have failed (Temussi
note that, based on previous research, the T1R3 subunit is required 2005). Three-dimensional structural modeling has revealed very lit-
for both umami and sweet modalities. If lactisole binds to the T1R3 tle similarity in the tertiary folds of the molecules (Tancredi and
subunit, it should inhibit both taste modalities. Lactisole indeed others 2004). The only observed structural similarities occur in the
inhibits both sweet and umami tastes in a dose-dependent manner secondary β-sheet loops; therefore, they have become the only
in humans. These findings were supported by mouse/human sweet fingers candidates in these proteins.
chimeras, also indicating that the T1R3 subunit is responsible for Tancredi and others (2004) designed several cyclic peptides cor-
binding lactisole (Jiang and others 2005b). Galindo-Cuspinera and responding to the “hairpin” structure of the secondary β-sheet
Breslin (2006) proposed that the T1R1 (umami) and T1R2 (sweet) loops of brazzein, monellin, and thaumatin; however, none elicited
subunits could be modulated by the T1R3 subunit and vice versa. 5 sweet taste, even at high concentrations. It may be that an incor-
Ribonucleotides prevent lactisole’s inhibition of umami taste, but rect primary structure (amino acid sequence) was chosen for the
not of sweet taste, suggesting that 5 ribonucleotides interact with peptides, so they lacked the residues necessary to generate a re-
the T1R1 umami receptor but not with the T1R2 sweet receptor. sponse, or that the secondary structure of the peptide lacked the
However, since it has been shown that lactisole binds to the T1R3 order necessary to fold into a similar 3-dimensional fold as the par-
receptor subunit, the 5 ribonucleotides’ interaction with the ent protein, or a combination of the two. However, the peptides
T1R1 unami receptor must affect the interaction of lactisole–T1R3 designed for this study exhibited the same hairpin conformations
subunit in some way. as their respective parent proteins. These researchers concluded
Some sweet compounds interact with other taste modalities, that these peptides were, in fact, an accurate model of the sweet
as well. Saccharin and acesulfame-K trigger both sweet and bit- residues identified as candidate sweet fingers on brazzein, mon-
ter taste modalities. These 2 ligands may cause bitter taste via a ellin, and thaumatin. The fact that the only observed similarity in
common mechanism that differs from other known bitter com- structure of these sweet proteins was the secondary β-sheet loops
pounds. Both saccharin and acesulfame-K activate 2 bitter recep- and peptides designed to mimic the sweet taste of these candidate
tors, hTAS2R43 and hTAS2R44, in transfected human kidney cells sweet fingers, and they were unable to elicit a sweet taste, suggests
(Kuhn and others 2004). These 2 receptors also exist in the taste that an alternate explanation is required.
papillae of the tongue, further suggesting that they interact with Spadacini and others (2003) reported an order of magnitude re-
saccharin and acesulfame-K. The sweet protein neoculin interacts duction in sweetness in a structural mutant of a single chained
with the sour taste modality (discussed subsequently; Shimizu- monellin, G16A MNEI, compared to its parent protein, MNEI.
Ibuka and others 2006; Koizumi and others 2007). Structural comparison of the G16A MNEI and its parent protein
showed only a slight rotation of the β-sheets with respect to the
helix. This rotation had a profound effect on the tertiary structure
Sweet protein binding: “sweet fingers”
of the molecule, but had no effect on the ability of the theorized
compared with the wedge model
sweet finger, Y65-D68, to interact with the sweet receptor. This sug-
Several sweet proteins can interact with the sweet taste receptor
gested that the area of interaction with the sweet receptor on the
(Temussi 2002, 2005, 2006a, 2006b; Spadacini and others 2003; Jiang
protein must include more surface area than the sweet finger Y65-
and others 2004, 2005a, 2005b; Tancredi and others 2004; Esposito
D68 alone, again calling into question the sweet fingers theory for
and others 2006; Shimizu-Ibuka and others 2006; Koizumi and oth-
sweet protein–receptor binding.
ers 2007). Proteins are unique in their receptor interaction in that
Using the crystal structure of mGluR1 as a homology model,
they are much larger molecules than the other ligands that bind to
docking brazzein into the closed cleft of hT1R2 allowed identifica-
the receptor; however, the same receptor binds all classes of sweet
tion of several residues at the mouth of the cleft that should interact
micro- and macromolecules.
with the sweet protein in the sweet fingers model. The disruptive
Three sweet proteins of known 3-dimensional structure appear
effect of mutations in these residues was examined in HEK cells.
frequently in the literature: brazzein, isolated from the African plant
Mutations at all 4 identified points of interaction had no effect on
species Pentadiplandra brazzeana; monellin, isolated from the
the brazzein–receptor interaction, suggesting that the sweet fingers
African plant species Dioscorephyllum cumminsii; and thaumatin,
model for brazzein is inaccurate. One of the mutations, D307N, re-
isolated from the African plant species Thaumatococcus daniel-
sulted in a complete loss of sweetness response when the receptor
lii (Kinghorn and Compadre 2001). These proteins have molecu-
interacted with small molecular weight sweeteners (D-tryptophan,
lar weights 15 to 65 times greater than that of sucrose, yet all 4
aspartame, sucrose; Jiang and others 2005a).
molecules dock to the same sweet receptor. A small component of
Using in silico 3-dimensional docking models, Temussi (2002)
the protein molecules’ structure may interact with the receptor in
proposed a secondary binding site on the T1R2–T1R3 receptor that
the same way as do small molecular weight sweeteners. Tancredi
is unique to sweet proteins. Of the more than 40 preferred solutions,
and others (2004) described this mechanism: small structures on
most centered on a large cavity of the T1R3 promoter. Additionally,
the surface of large proteins, termed “sweet fingers,” may exhibit
a strong candidate-binding site that could stabilize the sweet recep-
molecular structures similar to low molecular weight sweeteners
tor was the region that acts as a bridge between the T1R2 and T1R3
and can interact with the sweet receptor. These “sweet fingers,” by
promoter regions. However, this homology model construct con-
definition, would have to meet 3 criteria:
siders only 1 form of the receptor (T1R2 closed, T1R3 open; Wal-
1. They would have to be of sufficient length to interact with the ters and Hellekant 2006). Homology modeling of the T1R2 + T1R3
receptor. complex in both the aforementioned conformation and in the T1R2
2. Similar structures would have to be observed in different sweet open, T1R3 closed conformation showed brazzein to bind primar-
proteins. ily to T1R2 with contacts to T1R3 stabilizing the activated receptor
3. These similar structures would have to exhibit a similar shape or (Walters and Hellekant 2006).
sequence as the glycophores already identified in small sweeten- The crystal homology model of the N-terminal domain of the
ers. mGlu receptor can exist in 3 forms, a ligand-complexed (active)

Vol. 73, Nr. 6, 2008—JOURNAL OF FOOD SCIENCE R85


R: Concise Reviews Sweet taste review . . .
in Food Science

form and 2 free forms termed Free form I and Free form II. Free acterized with specific electrostatic changes in several neutrally
form II can exist in a conformation nearly identical to that of the charged residues. Docking calculations showed that the proposed
active complex. If the T1R2–T1R3 receptor behaves like the mGlu interaction surface of MNEI with the wedge site of the sweet re-
receptor, it would also exist as a mixture of these 3 forms; upon ceptor was positively charged. If monellin indeed interacts with
binding a small molecular weight ligand, the receptor in Free form the wedge site of the sweet receptor, changes in MNEI uncharged
I would stabilize to the active-conformation. Perhaps a 2nd mecha- residues into either acidic or basic states would affect the surface
nism for sweet receptor stabilization exists via a Free form I shift to that interacts with the sweet protein, and thus perceived sweet-
Free form II due to external binding of a macromolecule (for exam- ness of the molecule. Sensory studies, a more direct confirma-
ple, sweet protein; Tancredi and others 2004). Docking calculations tion of the validity of the Wedge Model for sweet protein–sweet
show that all 3 sweet proteins brazzein, monellin, and thaumatin fit receptor interaction than studies using human/mouse chimeras,
into a large cavity of the sweet receptor with wedge-shaped surfaces confirm this hypothesis. In addition, it offers insight into the im-
of their structure. This Wedge Model is currently the most widely portance of the electrostatic potential of sweet protein–receptor
accepted theoretical mechanism by which proteins can activate the interactions.
sweet receptor (Figure 5). Recently, the crystal structure of the sweet heterodimeric pro-
Jiang and others (2004) proposed an alternate binding site for tein neoculin, found in the fruit of Curculigo latifolia, was solved
sweet proteins located in the cysteine-rich region of the T1R3 re- (Shimizu-Ibuka and others 2006). Neoculin elicits a sweet response
ceptor. Using human/mouse chimeras of T1R3 paired with hu- in humans and also has taste-modifying activity in that it converts
man T1R2 (hT1R2), residues 536-545 in the cysteine-rich region sour taste to sweet taste. This is the 1st tertiary structure identi-
were required for a response to brazzein. Two amino acid substi- fied for a protein of this nature. Neoculin exists in pH-dependent
tutions (Ala-537 and Phe-540) eliminated all response. In addition, open and closed conformations. It may be that in the presence
mutations at the Phe-540 position reduced responses to brazzein of acid, the protein shifts into the open conformation, then in-
(and monellin), suggesting that, although brazzein and other sweet teracts with the T1R2–T1R3 receptor. This would account for the
proteins may interact with the T1R2 complex of the sweet recep- protein’s taste modifying activity. Using human/mouse chimeras
tor, interactions with the T1R3 complex are crucial to elicit a sweet of the T1R3 subunit, Koizumi and others (2007) demonstrated
response. It should be noted, however, that the validity of these that the NTD of human T1R3 (hT1R3) is required for the recep-
conclusions has been called into question; the same cysteine-rich tion of neoculin; a positive neoculin response occurred in hu-
region identified in this study has a predominant structural role in man T1R2–human T1R3 (hT1R2-hT1R3) cells, whereas a negative
other class C receptors (Esposito and others 2006). Therefore, mu- response occurred in human T1R2–mouse T1R3 (hT1R2-mT1R3)
tations in this region could undermine the structural integrity of cells.
the sweet receptor, causing the lack of sweet protein response as re-
ported by Jiang and others (2004). Walters and Hellekant (2006) also Sweet taste transduction
propose a greater likelihood that these mutations affect the confor- Transduction refers to the cascade of chemical signals occur-
mational change required to activate downstream 2nd messengers ring downstream from chemoreception that ultimately stimulates
as opposed to the actual binding of brazzein. the nervous system to signal the brain of an incoming stimulus.
Electrostatic potential is important in the sweet protein–sweet Although transduction pathways are generally similar for all senses,
receptor interaction in the Wedge Model (Esposito and others there are unique differences in the pathway of sweet taste, and
2006). Seven MNEI mutants were designed, expressed, and char- there is some debate as to the exact mechanism by which it occurs.
Prior to the discovery of the sweet taste receptor, several studies
showed changes in chemicals thought to be 2nd messengers (cAMP
and inositol 1,4,5-triphosphate [IP 3 ]) in signal transduction path-
Free Form I Complexed ways of TRCs in response to sweet ligands because these chemicals
A
increase in rats, hamster, and frog taste cells in response to a variety
Aspartame of sweet stimuli (Brand and Feign 1996; Kinnamon and Margolskee
1996). Increases in IP 3 in response to nonsaccharide ligands, but
not in response to sucrose, suggested that 2 different sweet taste
transduction mechanisms may exist for saccharide and nonsaccha-
ride sweeteners. A far greater increase in IP 3 concentration occurs
in response to the high potency sweeteners saccharin and SC-45647
than to sucrose, whereas a greater concentration of cAMP occurs in
Free Form I Free Form II response to sucrose (Bernhardt and others 1996).
B Margolskee (2002) proposed 2 models for sweet taste transduc-
Protein tion. In the first, a saccharide sweetener activates G s via the GPCR,
which activates adenyl cyclase (AC) generating cAMP. The cAMP
then acts directly, via an ion channel, or indirectly, via activation
of a protein kinase, to depolarize the taste cell via a release of
Ca2+ . Depolarization then triggers neurotransmitter release. The
2nd model describes a transduction pathway for nonsaccharide
Figure 5 --- Possible mechanisms of sweet molecule–
sweet receptor interactions (adapted from Tancredi and sweeteners. Upon binding to the GPCR, the nonsaccharide sweet-
others 2004). (A) Binding of aspartame transforms inac- ener activates phospholipase (PLCβ2) generating IP 3 and diacyl-
tive, ligand-free form I into the active complexed form. glycerol (DAG), which induce Ca2+ release from internal stores. As
(B) A sweet protein binding to a surface site of the ligand- in the 1st proposed transduction pathway, Ca2+ release causes cell
free form II shifts the equilibrium to favor active free form
II. Long-lasting signal transmission is due to stabilization depolarization and neurotransmitter release. Both mechanisms are
for form II by protein complexation. theorized to exist in the same TRC (Figure 6).

R86 JOURNAL OF FOOD SCIENCE—Vol. 73, Nr. 6, 2008


R: Concise Reviews
Sweet taste review . . .

in Food Science
Ca2+ and ion channels are crucial to the sweet taste transduction sweet and bitter compounds has been demonstrated in rat taste
pathway. Changes in Ca2+ concentration in TRCs occur in response bud cells in vivo showing that the tastants can physically interact
to several classes of sweeteners (Bernhardt and others 1996). The with the PKA and GRKs in taste cells (Zubare-Samuelov and others
mammalian transient receptor potential channel 5 (TRPM5), a 2005). The sweet tastants cyclamate, saccharin, and neohesperidin
cationic ion channel believed to regulate Ca2+ entry into cells, is dihydrochalcone (NHD), and several bitter compounds, inhibited
present in TRCs (Perez and others 2002, 2003). Several downstream GRK-phosphorylated rhodopsin and PKA-phosphorylated casein.
taste signaling molecules are coexpressed with TRPM5, supporting Interaction with protein kinases in the taste receptor might be the
the theory that this receptor regulates sweet (and/or bitter) trans- mechanism behind characteristic sweet and bitter aftertastes in a
duction. In mice, the absence of TRPM5 results in loss of sensitivity variety of compounds.
to saccharide and nonsaccharide sweeteners as well as bitter and
umami tastants (Zhang and others 2003). These observations high- Possible allosteric modulation of the sweet receptor
light the importance of Ca2+ and ion channel regulation for trans- The possibility that the class C GPCRs function allosterically has
duction of sweet taste and for bitter and umami tastes as well. been proposed (Pin and others 2005; Galindo-Cuspinera and oth-
Research by Sugimoto and others (2002) further supports the im- ers 2006). Different domains of the receptor exist in different con-
portance of ion channel activity in sweet taste transduction. Leptin, formations and may alter its activity. The NTD exists in both open
a hormone released by adipose tissue, inhibits food intake and in- and closed conformations, and in 2 orientations: resting “R” and
creases energy expenditure. In the taste cells of healthy mice, leptin active “A”. Oscillations between these conformations and orienta-
decreases sensitivity to sweet compounds by enhancing K+ influx tions provide potential allosteric modulation of the taste receptor
into the cell limiting sweet-induced depolarization. Genetically di- in response to ligand binding. This may explain unique sweetener
abetic db/db mice, which are deficient in the leptin receptor Ob- taste properties such as sweetener synergy and cross-adaptation
Rb gene, do not respond similarly to leptin. Sugimoto and others (Pin and others 2005).
(2002) argued that leptin’s interaction with Ob-Rb is essential to Galindo-Cuspinera and others (2006) pointed to an allosteric
K+ ion channel activation that inhibits sweet taste through hyper- model of the T1R1/T1R3 complex to explain the phenomenon of
polarization of the taste cell. sweet “water-taste.” Water-tastes are elicited by water after the
A recent hypothesis that nonsaccharide sweeteners interfere tastant chemical has been removed from the mouth; this effect
with downstream elements of the transduction pathway (Zubare- occurs for several gustatory sensations. To identify the mecha-
Samuelov and others 2005) is supported by characteristic sweet nism for this sensation in human sweet TRCs, hT1R1-hT1R3 cell
aftertaste of many of compounds (Schiffman and others 2007), sug- cultures have been used to replicate the action of rinsing with water
gesting possible interference with signal termination. G-protein re- in vitro. The known sweet water-taste stimuli, sodium saccharin,
ceptor signaling is reduced as a result of phosphorylation mediated and acesulfame-K, have been used to model the interaction. When
by 2 types of protein kinases: 2nd messenger-dependent protein ki- cells were exposed to 50 mM concentrations of the sweet com-
nases (PKA) and GPCR kinases (GRK). Phosphorylation inhibition pounds, a minimal response was observed; however, upon rins-
could delay signal termination. Membrane permeability of several ing, a strong response occurred. However, when exposed to 3 mM

A enyl
Adenyl GPCR GPCR
Cyclase

? ? ? ? Phospholipase ? ? ?

cAMP

IP 3
DAG
Protein
Kinase A
Ca 2
Ca 2
Na + Ca 2
Ca 2
Ca 2
K+
Ca 2
Ca 2

K+ K+ Na + K+ Ca 2

Ion Channels
Figure 6 --- Schematic representation of transduction mechanisms leading to taste cell depolarization (adapted from
Margolskee 2002).

Vol. 73, Nr. 6, 2008—JOURNAL OF FOOD SCIENCE R87


R: Concise Reviews Sweet taste review . . .
in Food Science

concentrations, the opposite effect was observed in that a strong question—What impact will this knowledge have on the future of
initial response occurred followed by no after-rinsing effect. The sweet taste? Will new, better tasting sweeteners and food products
authors suggest an allosteric model for the sweet taste receptor be developed? Will this research lead to a healthier population of
wherein high-affinity and low-affinity binding sites for saccharin adults in the world? A world in which people no longer enjoy sweet
and acesulfame-K occur. At low concentrations, these sweeten- tastes is hardly imaginable; therefore, it seems that the burden of
ers bind preferentially to the high-affinity site; at high concen- making it possible to continue to enjoy this pleasure while helping
trations, additional binding to the low-affinity site allostericallly to create a healthier population falls squarely on the shoulders of
shifts the receptor into an inactive conformation, inhibiting sweet food scientists. Further research in this field should provide food
taste. As the compounds are washed from the low-affinity binding scientists with the tools necessary to meet this challenge.
site, the receptor again assumes the active conformation and the
sweet taste returns. They concluded that sweet water-taste could Glossary of Terms
be used as predictor of sweet-taste blockers such as saccharin and Acesulfame-K: Acesulfame potassium is the potassium salt of a
acesulfame-K at high concentrations. dihydrooxathiazinone dioxide, which is approximately 200 times
sweeter than sugar (von Rymon Lipinski and Hanger 2001).
Effect of allelic variation of the T1R3 receptor Adenyl cyclase (AC): An enzyme involved in the manufacture of
Citing discrepancies between previously observed sweetener in- cAMP.
teractions in mice and rat species, an attempt to design a more Aspartame: A dipeptide sweetener composed of L-aspartic acid and
accurate system for measuring the sweetener response of the the methyl ester of L-phenylalanine, which is approximately 200
T1R3 subunit was undertaken by Bachmanov (2005). Many strains times sweeter than sugar (Butchko and others 2001).
of inbred, hybrid high sweetener-preferring and low sweetener- ATD: Amino terminal domain of the T1R2–T1R3 receptor complex,
preferring and 129.B6-Sac congenic mice were used because of also referred to as NTD.
the genetic variations in their T1R3 alleles. Varying responses to Brazzein: One of 2 sweet proteins found in the African plant species
a variety of sweet compounds were observed among the different Pentadiplandra brazzeana ranging from 500 to 2000 times sweeter
strains of mice, leading the author to conclude that allelic variation than sugar.
of the T1R3 gene affected response to some, but not all sweeten- CRD: Cysteine-rich domain of the T1R2–T1R3 receptor complex.
ers. An effect occurred in response to the sugars sucrose, glucose, Cross adaptation: The phenomenon of adaptation to a sweetener
maltose, and fructose; a sugar alcohol, erythritol; the amino acids resulting in the subsequent decreased sweet response to another
D-phenylalanine, D-tryptophan, and L-proline; and the artificial sweetener (Schiffman and others 2008).
sweeteners saccharin, acesulfame, sucralose, and SC45647, but not Cyclamate: Cyclamic acid often produced as a sodium or calcium
to the sweet amino acids glycine and L-alanine. Mice lacking or salt is a high potency sweetener approximately 30 times sweeter
expressing allelic variation in the T1R3 receptor genes responded than sugar (Bopp and Price 2001).
to several sweeteners (sucrose, saccharin, D-phenylalanine, D- Cyclic adenosine monophosphate (cAMP): A 2nd messenger in sig-
tryptophan, SC-45647, glycine, L-proline, L-alanine, L-glutamine), nal transduction that is a derivative of ATP and is manufactured by
indicating that many of these sweeteners interact with another the enzyme adenyl cyclase.
taste receptor, namely, the T1R2 subunit (Inoue and others 2004). Diacylglycerol (DAG): A 2nd messenger in the signal transduction
Allelic differences within species can result in observable pheno- pathway produced by the enzyme phopholipase.
typic differences in taste, which must be considered when evaluat- Dulcin: p-ethoxyphenylurea, a urea derivative, is a high potency
ing the sweet taste receptor. sweetener approximately 200 times sweeter than sugar (Kinghorn
and Compadre 2001).
Conclusions GA-1: Guanidinoacetic acid 1, a high potency sweetener estimated

T he accepted theory of the mechanism by which the human


sweet taste receptor functions has undergone many changes
in recent times. Early theories were inferentially based on the struc-
to be > 200000 sweeter than sugar when compared to 2% sucrose
(Nagarajan and others 1996).
G1-2: Guanidinoacetic acid 2, a high potency sweetener estimated
ture of the sweeteners, not on the receptor itself. With the discov- to be > 200000 sweeter than sugar when compared to 2% sucrose
ery of the T1R2/T1R3 receptor complex, understanding advanced (Nagarajan and others 1996).
as to how such a variety of sweet ligands, falling into such a wide Glycophore: The original term used by Schallenberger and Acree to
range of chemical classes, are able to bind to the same receptor. describe the theorized sweet receptor/sweet ligand complex.
Although most transduction pathways in humans show very sim- GPCR: G-protein coupled receptor.
ilar mechanisms, there are some mechanisms unique to the sweet HEK293: A strain of Human Embroytic Kidney Cells often used for
receptor. Although much has been gained in recent years, as with in vitro testing.
any new research, it seems that more questions have been created Heterodimer: A dimer composed of 2 different subunits.
than answers provided: Are the current homology models sufficient Inositol 1,4,5-triphosphate (IP 3 ): A 2nd messenger in signal trans-
for describing the human sweet receptor? Do sweeteners interfere duction made by the enzyme phopholipase C.
with downstream transduction signal pathways? Do bound ligands Lactisole: An aralkyl carboxylic acid that inhibits sweet and umami
change the shape of the sweet receptor allowing it to be allosteri- taste perception by humans but not by rats (Jiang and others
cally modified? How important is allelic variation? 2005b).
While work remains to be done, discovery of the sweet taste re- Ligand: A molecule that binds to a receptor.
ceptor and application of homology models allow for a far greater Monellin: The sweet protein found in the African plant species
understanding of sweet taste than was available only a few years Dioscorephyllum cumminsii ranging from 1500 to 2000 times
ago, making this a particularly exciting time for research in this sweeter than sugar.
field. It appears that it is only a matter of time until a complete mGluR1: A G-protein coupled glutamate receptor. The crystal
picture of the sweet taste pathway in humans from chemore- structure of the NTD has been solved and is often used as a homol-
ception through transduction is realized. However, this begs the ogy model for the sweet receptor.

R88 JOURNAL OF FOOD SCIENCE—Vol. 73, Nr. 6, 2008


R: Concise Reviews
Sweet taste review . . .

in Food Science
Neohesperidin dihydrochalcone (NHD): A glycosidic flavonoid that Freddolino PL, Yashar M, Kalani S, Vaidehi N, Floriano WB, Hall SE, Trabanino RJ, Kam
VWT, Goddard WA III. 2004. Predicted 3D structure for the human β2 adrenergic
is approximately 1800 times sweeter than sugar (Borrego and Mon- receptor and its binding site for agonists and antagonists. Proc Natl Acad Sci USA
tijano 2001). 101:2736–41.
Neotame: A structural derivative of the peptide sweetener aspar- Fredholm BB, Hökfelt T, Miligan G. 2007. G-protein-coupled receptors: an update.
Acta Physiol 190:3–7.
tame formed by alkylation of the molecule ranging from 7000 to Furse KE, Lybrand TP. 2003. Three-dimensional models for β-adrenergic receptor
13000 times sweeter than sugar (Stargel and others 2001). complexes with agonists and antagonists. J Med Chem 46:4450–62.
Galindo-Cuspinera V, Breslin PAS. 2006. The liaison of sweet and savory. Chem Senses
NTD: N-terminal domain of the T1R2-T1R3 receptor complex. 31:221–5.
Sac locus: The gene locus responsible for sweet taste originally Galindo-Cuspinera V, Winning M, Bufe B, Meyerhof W, Breslin PAS. 2006. A TAS1R
receptor-based explanation of sweet ‘water-taste.’ Nature 441:354–7.
named for the chemical compound saccharin. Goldsmith LA, Merkel CM. 2001. Sucralose. In: O’Brien Nabors L, editor. Alternative
Saccharin: Acid saccharin is a high potency sweetener approxi- sweeteners. 3rd ed., revised and expanded. New York: Marcel Dekker. p 185–207.
Goodman M, Del Valle JR, Amino Y, Benedetti E. 2002. Molecular basis of sweet taste
mately 300 times sweeter than sugar which is typically available as in dipeptide ligands. Pure Appl Chem 747:1109–16.
a sodium, calcium, or potassium salt (Pearson 2001). Gouldson PR, Kidley NJ, Bywater RP, Psaroudakis G, Brooks HD, Diaz C, Shire D,
Reynolds CA. 2004. Toward the active conformations of rhodopsin and the β 2 -
SC-45647: A guanidine-aliphatic acid several thousand times adrenergic receptor. Proteins 56:67–84.
sweeter than sugar (Walters and Hinds 1994). Hoon M, Adler E, Lindemeier J, Battey JF, Ryba NJP, Zuker CS. 1999. Putative mam-
malian taste receptors: a class of taste-specific GPCRs with distinct topographic
Sucralose: A high potency sweetener resulting from the chlorina- selectivity. Cell 96:541–51.
tion of sucrose in the 1 and 6 positions of the fructose moiety and Inoue M, Reed DR, Li X, Tordoff MG, Beauchamp GK, Bachmanov AA. 2004. Allelic
variation of the Tas1r3 taste receptor gene selectively affects behavioral and neural
the inversion and chlorination of the 4 position of the glucose moi- taste responses to sweeteners in the F 2 hybrids between C57BL/6ByJ and 129P3/J
ety that is approximately 600 times sweeter than sugar (Goldsmith Mice. J Neurosci 249:2296–303.
and Merkel 2001). 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
Sweetener synergy: The phenomenon of the combined sweetness 27943:45068–75.
of 2 or more sweeteners within a matrix being perceived as higher Jiang P, Cui M, Ji Q, Snyder L, Liu Z, Bernard L, Margolskee RF, Osman R, Max M.
2005a. Molecular mechanisms of sweet receptor function. Chem Senses 30(Suppl
than the sum of each individual sweetener. 1):i17–8.
T1R: A class of mammalian taste receptors for sweet and umami Jiang P, Cui M, Zhao B, Liu Z, Snyder LA, Benard LMJ, Osman R, Margolskee RF, Max
M. 2005b. Lactisole interacts with the transmembrane domains of human T1R3 to
taste modalities. inhibit sweet taste. J Biol Chem 28015:15238–46.
Thaumatin: A class of sweet proteins found in the arils of the fruits Jiang P, Cui M, Zhao B, Snyder LA, Benard LM, Osman R, Max M, Margolskee RF.
2005c. Identification of the cyclamate interaction site within the transmembrane
of the West African plant species Thaumatococcus daniellii ranging domain of the human sweet taste receptor subunit T1R3. J Biol Chem 28040:34296–
from 1600 to 3000 times sweeter than sugar (Kinghorn and Com- 305.
Kier LB. 1972. Molecular theory of sweet taste. J Pharmacy Sci 61:1394–7.
padre 2001). Kinghorn D, Compadre C. 2001. Less common high-potency sweeteners. In: O’Brien
TMD (7TM): 7-transmembrane domain of the T1R2–T1R3 receptor Nabors L, editor. Alternative sweeteners. 3rd ed., revised and expanded. New York:
Marcel Dekker. p 209–33.
complex. Kinnamon SC, Margolskee RF. 1996. Mechanisms of taste transduction. Curr Opin
TRC: Taste receptor cell. Neurobiol 6:506–13.
Kitagawa M, Kusakabe Y, Miura H, Ninomiya Y, Hino A. 2001. Molecular genetic iden-
Tumor necrosis factor receptor (TNFR): A family of receptors that tification of a candidate receptor gene for sweet taste. BioChem Biophys Res Comm
bind to a tumor necrosis factor. 2831:236–42.
VFTD: “Venus Flytrap” domain of the T1R2–T1R3 receptor complex Koizumi A, Nakajima K, Asakura T, Morita Y, Ito K, Shimizu-Ibuka A, Misaka T, Abe K.
2007. Taste-modifying sweet protein, neoculin, is received at human T1R3 amino
also referred to as NTD. terminal domain. Biochem Biophys Res Comm 358:585–9.
Kuhn C, Bufe B, Winnig M, Hofmann T, Frank O, Behrens M, Lewtschenko T, Slack JP,
Ward CD, Meyerhof W. 2004. Bitter taste receptors for saccharin and acesulfame-K.
References J Neurosci 244:10260–5.
Arterburn DE, Maciejewski ML, Tsevat J. 2005. Impact of morbid obesity on medical Kunishima N, Shimada Y, Tsuji Y, Sato T, Yamamoto M, Kumasaka T, Nakanishi S,
expenditures in adults. Int J Obes 29:334–9. Jingami H, Morikawa K. 2000. Structural basis of glutamate recognition by a dimeric
Bachmanov AA. 2005. Genetic approach to characterize interaction of sweeteners metabotropic glutamate receptor. Nature 407:971–7.
with the sweet taste receptors in vivo. Chem Senses 30(Suppl 1):i82–3. Li X, Inoue M, Reed DR, Huque T, Puchalski RB, Tordoff MG, Ninomiya Y, Beauchamp
Bernhardt SJ, Naim M, Zehavi U, Lindemann B. 1996. Changes in IP 3 and cytosolic GK, Bachmanov AA. 2001. High-resolution genetic mapping of the saccharin pref-
Ca2+ in response to sugars and non-sugar sweeteners in transduction of sweet taste erence locus Sac and the putative sweet taste receptor T1R1 gene Gpr70 to mouse
in the rat. J Physiol 4902:325–36. distal chromosome 4. Mammal Genome 12:13–6.
Bishay IE. 2006. The biology and chemistry of taste. Presented at the Illinois Inst. Tech- Li X, Staszewski L, Xu H, Durick K, Zoller M, Adler. 2002. Human receptors for sweet
nology, Chicago, Ill. and umami taste. Proc Natl Acad Sci USA 997:4692–6.
Bissantz C, Bernard P, Hibert M, Rognan D. 2003. Protein-based virtual screening of Margolskee RF. 2002. Molecular mechanisms of taste transduction. Pure Appl Chem
chemical databases. II. Are homology models of G-protein coupled receptors suit- 747:1125–33.
able targets? Proteins: Struct Function Genet 50:5–25. Max M, Shanker YG, Huang L, Rong M, Liu Z, Campagne F, Weinstein H, Damak S,
Bopp BA, Price P. 2001. Cyclamate. In: O’Brien Nabors L, editor. Alternative sweeten- Margolskee RF. 2001. Tas1r3, encoding a new candidate taste receptor, is allelic to
ers. 3rd ed., revised and expanded. New York: Marcel Dekker. p 63–85. the sweet responsiveness locus Sac. Nat Genet 281:58–63.
Borrego F, Montijano H. 2001. Neohesperidin dihydrochalcone. In: O’Brien Nabors Montmayeur JP, Liberles SD, Matsunami H, Buck LB. 2001. A candidate taste receptor
L, editor. Alternative sweeteners. 3rd ed., revised and expanded. New York: Marcel gene near a sweet taste locus. Nat Neurosci 4:492–8.
Dekker. p 87–104. Morini G, Bassoli A, Temussi PA. 2005. From small sweeteners to sweet proteins:
Brand JG, Feign AM. 1996. Biochemistry of sweet taste transduction. Food Chem anatomy of the binding sites of the human T1R2˙T1R3 receptor. J Med Chem
563:199–207. 48:5520–9.
Butchko HH, Stargel WW, Comer CP, Mayhew DA, Andress SE. 2001. Aspartame. In: Muto T, Tsuchiya D, Morikawa K, Jingami H. 2007. Structures of the extracellular re-
O’Brien Nabors L, editor. Alternative sweeteners. 3rd ed., revised and expanded. gions of the group II/III metabotropic glutamate receptors. Proc Natl Acad Sci USA
New York: Marcel Dekker. p 41–61. 104:3759–64.
Cherezov V, Rosenbaum DM, Hanson MA, Rasmussen SGF, Thian FS, Kobilka TS, Naim H, Huang L, Spielman AI, Shaul ME, Aliluiko A. 2006. Stimulation of taste cells
Choi HJ, Kuhn P, Weis WI, Kobilka BK, Stevens RC. 2007. High-resolution crystal by sweet taste compounds. In: Spillane WJ, editor. Optimizing sweet taste in foods.
structure of an engineered human β 2 -adrenergic G protein-coupled receptor. Sci- Boca Raton, Fla.: CRC Press. p 3–29.
ence 318:1258–65. Nagarajan S, Kellogg MS, DuBois GE. 1996. Understanding the mechanism of sweet
Cui M, Jiang P, Maillet E, Max M, Margolskee RF, Osman R. 2006. The heterodimeric taste: synthesis of ultrapotent guanidinoacetic acid photoaffinity labeling reagents.
sweet taste receptor has multiple potential ligand binding sites. Curr Pharmaceut J Med Chem 3921:4167–72.
Des 12:4591–600. Nelson G, Hoon MA, Chandrashekar J, Zhang Y, Ryba NJ, Zuker CS. 2001. Mammalian
Dubois GE. 2004. Unraveling the biochemistry of sweet and umami tastes. Proc Natl sweet taste receptors. Cell 106:381–90.
Acad Sci USA 10139:13972–3. Nie Y, Hobbs J Vigues S, Olson WJ, Conn CL, Munger SD. 2006. Expression and purifi-
Eggers SC, Acree TE, Schallenberger RS. 2000. Sweetness chemoreception theory and cation of functional ligand-binding domains of T1R3 taste receptors. Chem Sense
sweetness transduction. Food Chem 68:45–9. 31:505–13.
Esposito V, Gallucci R, Picone D, Saviano G, Tancredi T, Temussi TA. 2006. The impor- Nie Y, Vigues S, Hobbs JR, Conn GL, Munger SD. 2005. Distinct contributions of
tance of electrostatic potential in the interaction of sweet proteins with the sweet T1R2 and T1R3 taste receptor subunits to the detection of sweet stimuli. Curr Biol
taste receptor. J Mol Biol 360:448–56. 15:1948–52.
Flegal KM, Troiano RP. 2000. Changes in the distribution of body mass index of adults Nofre C, Tinti J. 1996. Sweetness reception in man: the multipoint attachment theory.
and children in the US population. Int J Obes 24:807–18. Food Chem 563:263–74.

Vol. 73, Nr. 6, 2008—JOURNAL OF FOOD SCIENCE R89


R: Concise Reviews Sweet taste review . . .
in Food Science

Ozeck M, Brust P, Xu H, Servant G. 2004. Receptors for bitter, sweet and umami taste Sugimoto K, Shigemura N, Yasumatsu K, Ohta R, Nakashima K, Kawai K, Ninomiya Y.
couple to inhibitory G protein signaling pathways. Eur J Pharm 489:139–49. 2002. Ion channels and second messengers involved in transduction and modula-
Pearson RL. 2001. Saccharin. In: O’Brien Nabors L, editor. Alternative sweeteners. 3rd tion of sweet taste in mouse taste cells. Pure Appl Chem 747:1141–51.
ed., revised and expanded. New York: Marcel Dekker. p 147–65. Tancredi T, Pastore A, Salvadori S, Esposito V, Temussi PA. 2004. Interaction of sweet
Perez CA, Huang L, Rong M, Kozak A, Preuss AK, Zhang H, Max M, Margolskee RF. proteins with their receptor: a conformational study of peptides corresponding to
2002. A transient receptor potential channel expressed in taste receptor cells. Nat loops of brazzein, monellin and thaumatin. Eur J Biochem 271:2231–40.
Neurosci 511:1169–76. Temussi PA. 2002. Why are sweet proteins sweet? Interaction of brazzein, monellin
Perez CA, Margolskee RF, Kinnamon SC, Ogura T. 2003. Making sense with TRP chan- and thaumatin with the T1R2-T1R3 receptor. Fed Exp Biol Soc Lett 526:1–4.
nels: store-operated calcium entry and the ion channel Trpm5 in taste receptor Temussi P. 2005. From oligopeptides to sweet proteins. J Pept Sci 111:262–4.
cells. Cell Calcium 33:541–9. Temussi P. 2006a. The history of sweet taste: not exactly a piece of cake. J Mol Recognit
Pin JP, Kniazeff J, Liu J, Binet V, Goudet C, Rondard P, Prezéau L. 2005. Allosteric 19:188–99.
functioning of dimeric class C G-protein-coupled receptors. Fed Exp Biol Soc Lett Temussi PA. 2006b. Natural sweet macromolecules: how sweet proteins work. Cell Mol
272:2947–55. Life Sci 63:1876–88.
Sainz E, Korley JN, Battey JF, Sullivan SL. 2001. Identification of a novel member of the von Rymon Lipinski GW, Hanger LY. 2001. Acesulfame-K. In: O’Brien Nabors L, editor.
T1R family of putative taste receptors. J Neurochem 77:896–903. Alternative sweeteners. 3rd ed., revised and expanded. New York: Marcel Dekker.
Salbe AD, DelParigi A, Pratley RE, Drewnowski A, Tataranni PA. 2004. Taste prefer- p 13–30.
ences and body weight changes in an obesity-prone population. Am J Clin Nut 79: Walters DE. 2002. Homology-based model of the extracellular domain of the taste re-
372–8. ceptor T1R3. Pure Appl Chem 74:1117–23.
Shallenberger RS. 1996. The AH-B glycophore and general taste chemistry. Food Walters DE, Hellekant G. 2006. Interactions of the sweet protein brazzein with the
Chem 563:209–14. sweet taste receptor. J Agric Food Chem 54:10129–33.
Schallenberger RS, Acree TE. 1967. Molecular theory of sweet taste. Nature 216: Walters DE, Hinds M. 1994. Genetically evolved receptor models: a computational
480–2. approach to construction of receptor models. J Med Chem 37:2527–36.
Schiffman SS, Sattely-Miller EA, Bishay IE. 2007. Time to maximum sweetness Xu H, Staszewski L, Tang H, Adler E, Zoller M, Li X. 2004. Different functional
intensity of binary and ternary blends of sweeteners. Food Qual Pref 18:405– roles of T1R subunits in the heteromeric taste receptors. Proc Nat Acad Sci USA
15. 10139:14258–63.
Schiffman SS, Sattely-Miller E, Bishay IE. 2008. Sensory properties of neotame: com- Yu L, Liang S, Liu X, He Q, Studholme DJ, Wu Q. 2004. NCD3G: a novel nine-cysteine
parison with other sweeteners. In: Weerasinghe DK, Dubois GE, editors. Sweetness domain in family 3 GPCRs. Trends Biochem Sci 29:458–61.
and sweeteners: biology, chemistry, and psychophysics. Washington, D.C.: Ameri- Zhao GQ, Zhang Y, Hoon MA, Chandrashekar J, Erlenbach I, Ryba NJ, Zuker CS. 2003.
can Chemical Society. p 511–29. The receptors for mammalian sweet and umami taste. Cell 115:255–66.
Shimizu-Ibuka A, Morita Y, Terada T, Asakura T, Nakajima K, Iwata S, Misaka T, Zhang Y, Hoon M, Chandrashekar J, Mueller KL, Cook B, Wu D, Zuker CS, Ryba NJP.
Sorimachi H, Arai S, Abe K. 2006. Crystal structure of neoculin: insights into its 2003. Coding of sweet bitter and umami tastes: different receptor cells sharing sim-
sweetness and taste-modifying activity. J Mol Biol 359:148–58. ilar signaling pathways. Cell 112:293–301.
Spadacini R, Trabucco F, Saviano G, Picone D, Crescenzi O, Tancredi T, Temussi Zhang Y, Devries ME, Skolnick J. 2006. Structure modeling of all identified G protein–
PA. 2003. The mechanism of interaction of sweet proteins with the T1R2-T1R3 re- coupled receptors in the human genome. PLoS Comput Biol 2:88–99.
ceptor: evidence from the solution structure of G16A-MNEI. J Mol Biol 328:683– Zubare-Samuelov M, Shaul ME, Peri I, Aliluiko A, Tirosh O, Naim M. 2005. Inhibition
92. of signal termination-related kinases by membrane-permeant bitter and sweet tas-
Stargel WW, Mayhew DA, Comer CP, Andress SE, Butchko HH. 2001. Neotame. In: tants: potential role in taste signal termination. Am J Physiol Cell Physiol 289:483–
O’Brien Nabors L, editor. Alternative sweeteners. 3rd ed., revised and expanded. 92.
New York: Marcel Dekker. p 129–45.

R90 JOURNAL OF FOOD SCIENCE—Vol. 73, Nr. 6, 2008

You might also like