Está en la página 1de 12

OMICS A Journal of Integrative Biology

Volume 13, Number 1, 2009


© Mary Ann Liebert, Inc.
DOI: 10.1089/omi.2008.0031

Review

Genetic Variation in Taste and Its Influence


on Food Selection

Bibiana Garcia-Bailo, Clare Toguri, Karen M. Eny, and Ahmed El-Sohemy

Abstract

Taste perception plays a key role in determining individual food preferences and dietary habits. Individual dif-
ferences in bitter, sweet, umami, sour, or salty taste perception may influence dietary habits, affecting nutri-
tional status and nutrition-related chronic disease risk. In addition to these traditional taste modalities there is
growing evidence that “fat taste” may represent a sixth modality. Several taste receptors have been identified
within taste cell membranes on the surface of the tongue, and they include the T2R family of bitter taste re-
ceptors, the T1R receptors associated with sweet and umami taste perception, the ion channels PKD1L3 and
PKD2L1 linked to sour taste, and the integral membrane protein CD36, which is a putative “fat taste” recep-
tor. Additionally, epithelial sodium channels and a vanilloid receptor, TRPV1, may account for salty taste per-
ception. Common polymorphisms in genes involved in taste perception may account for some of the in-
terindividual differences in food preferences and dietary habits within and between populations. This variability
could affect food choices and dietary habits, which may influence nutritional and health status and the risk of
chronic disease. This review will summarize the present state of knowledge of the genetic variation in taste,
and how such variation might influence food intake behaviors.

Introduction taste modality (Laugerette et al., 2005). Perception of each of


these taste modalities appears to be mediated by a different

I NDIVIDUALS MAKE FOOD CHOICES based on a number of phys-


iological, nutritional, environmental, and sociocultural
factors. However, the sensory qualities of food are critical to
mechanism, although the molecular basis for some of them
remains relatively obscure. Ion channels in TRCs account for
salty taste perception (Drayna, 2005), while different G-pro-
dietary preferences, and taste in particular may be the most tein coupled receptors bind bitter (Adler et al., 2000; Chan-
important determinant of food choices (Drewnowski and drashekar et al., 2000; Matsunami et al., 2000), sweet (Li et
Rock, 1995; Glanz et al., 1998; Leterme et al., 2008). Taste per- al., 2002), and umami tastants (Nelson et al., 2002). Sour tas-
ception occurs when chemical molecules from food reach mi- tants are detected by members of the transient receptor po-
crovilli located at the apical tip of taste receptor cells (TRCs) tential ion channel family (Huang et al., 2006; Ishimaru et al.,
(Chandrashekar et al., 2006; Ishimaru et al., 2006; Linde- 2006; LopezJimenez et al., 2006). The genes encoding several
mann, 2001). TRCs congregate in groups of 50 to 100 to form of these receptors have been identified, and they include the
onion-shaped taste buds (Bachmanov and Beauchamp, 2007; TAS2R gene family for bitter (Adler et al., 2000; Chan-
Ishimaru et al., 2006; Lindemann, 2001). Taste buds, in turn, drashekar et al., 2000; Matsunami et al., 2000), the TAS1R
cluster into circumvallate, foliate, or fungiform papillae, and family for sweet and umami (Bachmanov et al., 2002; Li et
mammalian papillae are located in epithelial surfaces of the al., 2002), and PKD2L1 and PKD1L3 for sour taste (Huang et
tongue, palate, pharynx, larynx, and upper esophagus al., 2006; Ishimaru et al., 2006; LopezJimenez et al., 2006). In
(Chandrashekar et al., 2006; Kataoka et al., 2008; Lindemann, addition, the fatty acid transporter CD36 has been recently
2001). identified as a putative taste receptor for fat (Laugerette et
The vast array of flavors, defined as overall sensory per- al., 2005).
ception, found in foods is triggered by only a few, distinct Genetic variation in taste receptors may account for dif-
taste modalities: bitter, sweet, salty, sour, and umami (sa- ferences in food choices and dietary habits. Polymorphisms
vory) (Bachmanov and Beauchamp, 2007; Roper, 2007). of the genes that code for taste receptors may explain some
There is growing evidence that “fat” may also be a distinct of the observed variability in taste perception. This variabil-

Department of Nutritional Sciences, Faculty of Medicine, University of Toronto, Toronto, Ontario, M5S 3E2, Canada.

69
70 GARCIA-BAILO ET AL.

ity could affect food choices and dietary habits, and might and PROP are often used in taste perception studies to as-
influence nutritional and health status, as well as the risk of sess the link between genetics and bitter taste perception. A
chronic disease (Fig. 1). This review will provide a summary recent review covers PTC and PROP tasting study method-
of the current knowledge of taste perception mechanisms ologies and evolutionary implications of these phenotypes,
and genetic variation in taste, and how such variation might and offers an in-depth summary of their association with
influence food intake behaviors. food acceptance (Tepper, 2008).
Overall, about 75% of humans find PTC and PROP bitter
in taste, and they are classified as “tasters.” The remainder
Bitter Taste
of the population finds PTC and PROP tasteless and are con-
Bitter has been the most extensively studied taste modal- sidered “nontasters” (Kim and Drayna, 2004). The propor-
ity. Bitter tasting compounds are ubiquitous in nature and tion of nontasters varies between populations across the
structurally diverse at the molecular level (Chandrashekar globe. In West Africa, 3% of the population is blind to the
et al., 2006; Kim et al., 2004). Many bitter tasting substances bitter taste of PTC and PROP, but in China the proportion
are noxious, and bitter taste perception probably evolved to ranges between 6% and 23%, and in India it has been re-
prevent the consumption of plant toxins (Chandrashekar et ported to be as high as 40%. Caucasian North American pop-
al., 2006; Drewnowski and Rock, 1995; Kim and Drayna, ulations exhibit a nontaster frequency of about 30% (Guo and
2004). Bitter stimuli elicit rejection responses in young chil- Reed, 2001; Tepper, 1998). While the distribution of sensi-
dren and nonhuman primates (Roper, 2007). However, di- tivity to PTC/PROP bitterness was traditionally described
etary sources of bitter tastants are common, and they include as bimodal, increasing evidence suggests that the trait dis-
nutritionally significant plants such as spinach, endives, and plays a broad, continuous distribution (Bartoshuk et al., 1994;
many cruciferous vegetables, such as broccoli, bok choy, Kim and Drayna, 2004). Within the portion of the popula-
kale, cauliflower, cabbage, watercress, and arugula. Other tion that can taste PTC/PROP, there exists variability in the
bitter-tasting foods include sharp cheeses, soy products, perceived degree of bitterness such that tasters can be sub-
grapefruit, beer, green tea and coffee (Anliker et al., 1991; divided into medium tasters and supertasters (Bartoshuk et
Drewnowski et al., 2001; Duffy and Bartoshuk, 2000; Kamin- al., 1994; El-Sohemy et al., 2007). Additionally, increased sen-
ski et al., 2000; Keller et al., 2002; Prescott et al., 2004; Turn- sitivity to bitter taste has been hypothesized to correlate with
bull and Matisoo-Smith, 2002). These foods contain bitter- greater general taste acuity. A number of studies have found
tasting phytochemicals such as isothiocyanates resulting associations between increased bitterness sensitivity and
from glucosinolate breakdown in cruciferous vegetables, as heightened responses to sweet (Bartoshuk et al., 1994; De-
well as polyphenols, methylxanthines, isoflavones, and sul- Simone and Lyall, 2006; Duffy et al., 2004; Mennella et al.,
famides. All of these phytochemicals might protect against 2005) and salty stimuli (Bartoshuk et al., 1998). Oral irritants
a number of illnesses (Basson et al., 2005; Drewnowski and such as capsaicin and ethanol also elicit heightened re-
Rock, 1995; El-Sohemy et al., 2007; Timpson et al., 2005). In- sponses among PTC/PROP tasters (Prescott and Swain-
dividuals who perceive these compounds as more intensely Campbell, 2000). Furthermore, tasters show increased sensi-
bitter may avoid consuming them, and this could affect their tivity to olfactory cues (Pickering et al., 2006) and to viscous
nutritional and health status. substances such as fats and food thickeners (Bartoshuk,
Bitter taste perception is a variable trait, and its genetic ba- 2000).
sis was identified over 75 years ago through a series of stud- Research conducted in the last decade has shown that bit-
ies on individual responses to phenylthiocarbamide (PTC) ter taste perception is mediated by the T2Rs, a family of G
(Blakeslee, 1932; Fox, 1931). PTC and the related compound protein-coupled taste receptors with seven trans-membrane
6-n-propylthiouracil (PROP) are members of the thioureas domains and a short extracellular N-terminus (Bachmanov
and contain a thiocyanate moiety (N-C  S) (Bartoshuk et and Beauchamp, 2007; Roper, 2007). T2Rs are located at the
al., 1994; Tepper, 1998). While PTC and PROP do not occur surface of taste cells within circumvallate and foliate papil-
naturally in foods, they elicit a bitter taste stimulus in some lae, and to a lesser extent within fungiform papillae. They
individuals. The variability in response to their taste corre- are also present in the palate and epiglottis (Adler et al., 2000;
lates with taste sensitivity to other bitter substances present Bachmanov and Beauchamp, 2007; Chandrashekar et al.,
in foods (Barnicot et al., 1951; Blakeslee and Salmon, 1935; 2000; Matsunami et al., 2000). In humans, T2Rs are encoded
Drewnowski and Rock, 1995; Hall et al., 1975). Thus, PTC by about 25 functional TAS2R genes located on chromo-

FIG. 1. Gene association studies have linked genetic variation in taste receptors to risk of disease. This can occur through
differences in taste perception, which may lead to differences in food preferences and food intake. This variation in food
intake may, in turn, affect metabolism and health outcomes.
GENETIC VARIATION IN TASTE 71

somes 5, 7, and 12 (Drayna, 2005; Tepper, 2008). Bachmanov nontasters, consumed more cruciferous vegetables than in-
and Beauchamp (2007) provide a thorough review of the dis- dividuals carrying a copy of the PA haplotype. The AV and
covery and genomic organization of these and other taste re- PA haplotypes were used as equivalent to AVI and PAV, re-
ceptor genes. spectively.
Over the past several years, efforts have been made to link The effects of TAS2R38 genetic variation on dietary pref-
T2R bitter taste receptors to particular tastants (Bufe et al., erences and subsequent coronary heart disease (CHD) risk
2002; Kuhn et al., 2004). In particular, TAS2R38 has been were examined in postmenopausal women (Timpson et al.,
identified as the gene for PTC sensitivity (Kim et al., 2003). 2005). No significant associations were found between geno-
TAS2R38, located on chromosome 7q, consists of a single type, CHD, or diet. However, the nontaster genotype,
1002-bp long exon that encodes a 333-amino acid, 7 trans- AVI/AVI, was associated with a slightly lower risk of dia-
membrane-spanning domain, guanine nucleotide-binding betes, suggesting that these women may have consumed a
protein-coupled receptor (Drayna, 2005). Three functional more prudent diet, perhaps consisting of more bitter-tasting
single nucleotide polymorphisms (SNPs) within this gene vegetables, over their lifetime. No direct link was found be-
have been found to explain up to 85% of the observed vari- tween TAS2R38 genotype and dietary habits, but it is possi-
ance in PTC taste sensitivity (Drayna, 2005; Kim et al., 2003). ble that an association between these two variables was
These polymorphisms encode for amino acid substitutions masked by the age of the subjects, because taste acuity is
at position 49 (alanine/proline, A49P), 262 (valine/alanine, known to diminish with age (Navarro-Allende et al., in press;
V262A), and 296 (isoleucine/valine, I296V) (Drayna, 2005; Whissell-Buechy, 1990). Interestingly, TAS2R38 taster geno-
Kim et al., 2003). While several resulting haplotypes have types have been observed to associate with a preference for
been observed, the two most common are PAV and AVI, sweet-tasting foods in children, but not in adults (Mennella
which correspond to the amino acids at the three positions. et al., 2005).
PAV homozygotes exhibit the greatest sensitivity to PTC/ Other members of the T2R gene family are involved in de-
PROP bitterness, while AVI homozygotes are the least sen- tecting bitter taste, and may influence dietary habits. For ex-
sitive (Bufe et al., 2005; El-Sohemy et al., 2007; Kim and ample, a single nucleotide polymorphism in TAS2R16, which
Drayna, 2004). Heterozygotes (PAV/AVI) show intermediate codes for a taste receptor for bitter -glucopyranosides, has
bitter taste sensitivity. Interestingly, some research indicates been observed to associate with alcohol dependence (Hin-
that TAS2R38 polymorphisms are not sufficient to explain richs et al., 2006; Wang et al., 2007). A possible link between
PROP bitterness perception at suprathreshold concentra- a polymorphism in TAS2R50 and increased risk of myocar-
tions, suggesting that other genetic or environmental mech- dial infarction has been identified as well (Shiffman et al.,
anisms may also play a role in PROP taste perception (Bufe 2005, 2008). This same polymorphism might be associated
et al., 2005; Hayes et al., 2008). with poor dietary habits. However, no studies to date have
It has been hypothesized that individuals with increased examined the potential relationship between TAS2R50 geno-
bitter taste sensitivity might avoid antioxidant-rich vegeta- type and food preferences or dietary habits.
bles because of their perceived bitterness, consuming instead
sweet, fatty foods and potentially increasing their risk of car-
Sweet Taste
diovascular and other diseases. However, an increased sen-
sitivity to bitterness has also been associated with height- Unlike bitter taste perception, sweet substances are per-
ened taste acuity (Bartoshuk et al., 1994, 1998; Chang et al., ceived as pleasant and are clustered separately from bitter
2006; Duffy, 2004), which may prevent food overconsump- taste in humans, possibly reflecting evolutionary pressures
tion in general. A number of associations have been found to select foods high in energy (Hladik et al., 2002). It is well
between perceived PTC or PROP bitterness and food pref- established that sweet tasting substances induce cephalic-
erences. For example, in young women there is an inverse phase reflexes, and therefore, sweet taste receptors in the
relationship between PROP sensitivity and acceptance of tart tongue and palate may be important in initiating the pre-ab-
citrus, Brassica vegetables, spinach, and coffee (Drewnowski sorptive metabolic response to food consumption (Tordoff,
et al., 1998, 1999). Female PROP tasters also show a decreased 1988; Zafra et al., 2006). In addition to natural sugars such
acceptance of sweet and fatty foods (Duffy and Bartoshuk, as glucose, fructose, sucrose, and sugar alcohols such as sor-
2000). bitol, a variety of artificial sweeteners have been developed
While there seems to be an association between bitterness (Knight, 1994; Mazur, 1984; Parker, 1978). Examples include
sensitivity and food preferences, the potential interaction be- saccharin, acesulfame-K, aspartame, sucralose, and dulcin
tween bitterness sensitivity and actual food consumption has (Boughter and Bachmanov, 2007). Additionally, amino acids
yet to be fully understood. In a group of young women, per- such as glycine, D-phenylalanine, D-tryptophan, L-proline,
ceived PROP bitterness was not associated with consump- and L-glutamine also confer a sweet taste (Boughter and
tion frequency of 22 bitter foods (Kaminski et al., 2000). How- Bachmanov, 2007). However, it is not clear why organisms
ever, among men undergoing endoscopic screenings for have evolved to detect these nonnutritive sweeteners. De-
colon polyps, those with the highest PROP sensitivity ate spite the current use of artificial sweeteners to reduce calo-
fewer vegetables (Basson et al., 2005). Interestingly, per- ries, there is evidence that noncaloric sweeteners may pro-
ceived PROP bitterness also correlated positively with polyp mote obesity by interfering with the normal metabolic
number, suggesting a possible link between PROP sensitiv- response to food intake (Swithers and Davidson, 2008).
ity and colon cancer. A similar link between increased bit- Using differences in sweet taste perception of saccharin
terness sensitivity and decreased vegetable consumption among different strains of mice, Fuller began to pave the way
was found in an Italian population (Sacerdote et al., 2007). for the search to identify the Sac locus (Fuller, 1974). The dis-
AV/AV homozygotes for TAS2R38, who were considered covery of the gene responsible for the Sac locus was greatly
72 GARCIA-BAILO ET AL.

accelerated 25 years later after a novel G-protein coupled re- counted for a greater genetic contribution to sweet taste dis-
ceptor family, TR1, was discovered in rats and humans crimination (Keskitalo et al., 2007a). Similarly, among related
(Hoon et al., 1999). In 2001, several groups identified the and unrelated family members, heritability estimates for de-
third member of the taste receptor family 1, T1R3 (TAS1R3 tecting intensity of suprathreshold sucrose solutions were
in humans, or Tas1r3 in rodents) as the gene responsible for near zero, yet the heritability of pleasantness of the two
the saccharin preferring phenotype, which was discovered strongest solutions and of sweet foods were between 30 and
using a number of different approaches (Bachmanov et al., 40% (Keskitalo et al., 2007b). This study identified a quanti-
2001; Kitagawa et al., 2001; Max et al., 2001; Montmayeur et tative trait linkage for use frequency of sweet foods on chro-
al., 2001; Nelson et al., 2001; Sainz et al., 2001). Direct evi- mosome 16p11.2, but no candidate genes were identified be-
dence for the role of Tas1r3 in saccharin preference came cause there are no genes known to affect sweet preference
from a study that isolated Tas1r3 from taster mice and res- and use in this region (Keskitalo et al., 2007b). The variabil-
cued nontaster mice by transfecting them with the taster ity in heritability observed in both studies, however, dem-
form of the gene (Nelson et al., 2001). In addition to the onstrates that preference for sweet foods is a multifactorial,
Tas1r3 gene, these studies also identified the coexpression of polygenic trait (Keskitalo et al., 2007a, 2007b). This is not sur-
T1R2 and T1R3 in circumvallate and foliate cells (Max et al., prising given the complexity of eating behaviors, which are
2001; Montmayeur et al., 2001; Nelson et al., 2001) and us- influenced by other physiological systems including food re-
ing a heterologous assay system, it was discovered that the ward circuits and energy homeostasis in addition to differ-
T1R2/T1R3 heteromer responded to a number of sweet sub- ences in taste perception. We have previously demonstrated
stances as measured using a calcium indicator dye (Nelson that a genetic variant in the glucose transporter type 2
et al., 2001). It was later discovered that T1R3 in combina- (GLUT2) gene is associated with a higher habitual con-
tion with T1R1 is the heteromer responsible for umami taste sumption of sugars, which suggests that glucose-sensing
detection (Nelson et al., 2002). Alpha-gustducin, which is a mechanisms that may be involved in glucose homeostasis
G-protein coupled receptor subunit, is proposed to be one and/or energy balance may affect food intake behaviours
pathway involved in transmitting the signal downstream (Eny et al., 2008). Furthermore, these different biological
once the sweet taste receptor is activated by its ligand (Mar- pathways may converge and interact to influence the over-
golskee, 2002; Sainz et al., 2007). Bitter and umami taste re- all consumption of sugars. Interestingly, T1R2 and T1R3
ceptors could also transmit their signal via alpha-gustducin have been localized in the small intestine and have been im-
(Margolskee, 2002; Sainz et al., 2007). Thus, in addition to plicated in increasing expression of GLUT2 (Mace et al., 2007)
understanding taste receptors, understanding the down- and the sodium-dependent glucose transporter isoform 1
stream transduction pathway is key to obtaining a complete (SGLT1) (Margolskee et al., 2007) in the brush-border mem-
picture of taste perception. brane in response to natural and artificial sweeteners in rats
Despite the discovery of the T1R2/T1R3 heteromer re- and mice (Mace et al., 2007; Margolskee et al., 2007). Thus,
sponsible for sweet taste detection, the number of sweet taste examining polymorphisms within the sweet taste receptor
receptors that exist is still unresolved. Evidence from a sin- genes may contribute to understanding part of the differ-
gle and double knockout of Tas1r3 and Tas1r2 suggests that ences in sweet taste detection and liking of sugars as well as
T1R2 and/or T1R3 may act as a low affinity taste monomer differences in postprandial glucose absorption, which may
or homodimer under high concentrations of natural sugars help explain predispositions to obesity, diabetes, and other
(Zhao et al., 2003), while another Tas1r3 knockout study sug- eating disorders.
gests that other sweet taste receptors might exist (Delay et The human gene encoding T1R3, TAS1R3, is located on
al., 2006). Interestingly, members of the Felidae family of ob- chromosome 1 together with the first and second members
ligate carnivores are unique because they are indifferent to of the taste receptor family 1, TAS1R1 and TAS1R2, respec-
sweet tasting foods and do not show neural responses to sug- tively (Liao and Schultz, 2003). Several polymorphisms have
ars (Li et al., 2005). This phenotype was explained by a mi- been identified in each of these three taste receptor genes,
crodeletion and early stop codon in the Tas1r2 gene, result- which varied across the different populations studied in-
ing in lack of T1R2 expression in taste tissue, thereby cluding Asian, Native American, African, and European
highlighting the importance of T1R2 in sweet taste detection. populations (Kim et al., 2006). A majority of the nonsyn-
Future studies are needed to resolve the debate regarding onymous variations reside in the extracellular domain of the
the number of proteins acting as a sweet taste receptor us- protein, which is hypothesized to contain the ligand-bind-
ing both Tas1r3 and Tas1r2 knockouts alone and in combi- ing site for carbohydrates and dipeptide sweeteners (Kim et
nation. al., 2006). TAS1R2 notably had higher variation than TAS1R1
Within humans, it has long been recognized that there are and TAS1R3, and was determined to be in the top 10th per-
interindividual differences in sweet taste detection thresh- centile of genetic diversity in comparison to 3305 other genes
olds (Blakeslee and Salmon, 1935; Henkin and Shallenberger, (Kim et al., 2006). Together with the remarkable genetic di-
1970). A recent study of female monozygotic and dizygotic versity, TAS1R2 was also determined to have potentially
twins reported that the additive genetic contribution to the evolved to sense a wide variety of structurally different
discrimination of the intensity of a sweet solution was 33%, sweet substances as assessed by rejecting the evolutionary
while the additive genetic contribution to liking the sweet neutrality test, called Tajima’s D test (Kim et al., 2006). No
solution was 49% (Keskitalo et al., 2007a). The solution used study has yet examined whether genetic variants in TAS1R2
in this study was a very sweet 20% sucrose solution and, and/or TAS1R3 are associated with differences in the con-
therefore, a lower sucrose concentration may have had the sumption of sugars in humans. Together with chimera stud-
capacity to further identify interindividual differences in ies evaluating the ligand binding and active sites of the het-
sweet taste discrimination, which may have potentially ac- eromeric taste receptor (Nie et al., 2005; Xu et al., 2004),
GENETIC VARIATION IN TASTE 73

candidate polymorphisms may be investigated to begin to enhancer IMP. Furthermore, cells transfected with human
understand the interindividual differences in sweet taste per- T1R1/T1R3 responded much more strongly to glutamate, the
ception and consumption of sugars. main human umami tastant, than cells expressing mouse
T1R1/T1R3 did. This observation was attributed to taste re-
ceptor sequence differences between the two species, per-
Umami Taste
haps reflecting differences in their natural diets. Overall,
The word umami is used to describe meaty, savory flavor, these results suggested a role for T1R1/T1R3 as a broad L-
and it comes from a Japanese term meaning “good taste” or amino acid receptor, which responds more specifically to
“delicious” (Chandrashekar et al., 2006; Roper, 2007). As is various taste stimuli depending on the animal species being
the case with sweet taste, animals are attracted to umami studied (Nelson et al., 2002).
taste. The main substance eliciting umami taste in humans Further supporting the role of T1R1/T1R3 as an umami
is L-glutamate, an amino acid abundantly found in food that taste receptor, it was shown that human umami tastants
often occurs as monosodium glutamate (MSG). L-aspartate caused excitation of HEK cells only when the cells coex-
also elicits umami taste in humans (Chandrashekar et al., pressed human T1R1 and T1R3 (Li et al., 2002). The response
2006; Zhao et al., 2003). The umami taste of MSG and L-as- was greater in the presence of IMP or GMP, and T1R1/T1R3
partate are greatly enhanced by the purine nucleotides ino- did not respond to sweet stimuli, D-amino acids, or either
sine 5-monophosphate (IMP) and guanosine 5-monophos- IMP or GMP alone. Finally, Tas1R1 and Tas1R3 knockout
phate (GMP) (Chandrashekar et al., 2006; Kurihara and mice were characterized both through behavioural tests and
Kashiwayanagi, 2000; Nelson et al., 2002). Umami tastants by measuring activity of the gustatory chorda tympani nerve
are found naturally in a wide array of vegetables such as after exposure to different taste stimuli (Zhao et al., 2003).
tomatoes, potatoes, mushrooms, carrots, and various sea- As expected, Tas1R1 and Tas1R3 knockout mice showed a
weeds, as well as fish, seafood, meat, and cheese (Kurihara complete loss of preference for umami, and they exhibited
and Kashiwayanagi, 2000). no chorda tympani nerve activity after stimulation with glu-
Umami was first described by Ikeda in 1909. The original tamate. Thus, it seems clear that umami taste is processed
description, published in Japanese, was only translated into by the heteromeric T1R1/T1R3 receptor (Li et al., 2002; Nel-
English in 2002 (Ikeda, 2002). Widespread acceptance of son et al., 2002; Zhao et al., 2003).
umami as a distinct taste modality occurred slowly. This was To date, human umami taste perception variability re-
partly because umami substances have a subtle taste even at mains poorly understood. Individual responses of European
high concentrations, and partly because the umami taste of adults to L-glutamate were assessed through detection
L-glutamate can be difficult to dissociate from the salty taste threshold determination, suprathreshold evaluation, per-
of the sodium also found in MSG (Kim et al., 2004; Linde- ception of stimulus strength and duration, and qualitative
mann et al., 2002). However, in the past decade several assessment of taste of MSG versus NaCl (Lugaz et al., 2002).
umami taste receptors have been proposed. Chaudhari et al. About 27% of subjects assessed were unable to distinguish
(1996, 2000) first described a modified brain glutamate re- MSG from NaCl at 29 mM isomolar concentrations, mean-
ceptor, the metabotropic G-protein receptor taste-mGluR4, ing they were unable to separate the umami taste compo-
which has a truncated N-terminus and is expressed in rat nent from the salty component of MSG. In time-intensity
taste buds. This receptor elicited identical behavioural re- tests of MSG versus NaCl stimuli, 20% of the participants
sponses in rats when stimulated by either MSG or L-2-amino- had a weakened MSG stimulus response. These results sug-
4-phosphonobutyrate (L-AP4), a ligand for mGluR4 that also gest a reduced ability to taste umami, but no attempt was
elicits umami taste responses in humans. Furthermore, Chi- made to address the genetic basis of the observed MSG in-
nese hamster ovary cells expressing taste-mGluR4 re- sensitivity, and the heritability of this trait has yet to be as-
sponded to stimulation by L-AP4 and MSG in concentrations certained in family or twin studies (Kim et al., 2004).
similar to those necessary to elicit taste responses in vivo. Complete DNA sequences of the TAS1R gene family cod-
These results were interpreted as evidence for the role of ing regions were compared across Asian, Native American,
taste-mGluR4 in umami taste perception. However, addi- African, and European populations (Kim et al., 2006). Sev-
tional evidence suggests that receptors other than taste- eral SNPs were identified within the extracellular domain of
mGluR4 may play a more important role in umami taste per- TAS1R1 and TAS1R3, and their frequencies varied between
ception. Most notably, the taste-mGluR4 receptor lacks a populations. This evidence might suggest interindividual
large portion of the domain necessary for glutamate recog- variability in umami taste perception, but the relationship
nition, and mGluR4 knockout mice still respond to umami between genetic polymorphisms in TAS1R1 or TAS1R3,
stimuli (Chaudhari and Roper, 1998; Zhao et al., 2003). umami taste perception and preference for foods with this
Umami processing seems to be closely related to sweet taste has yet to be explored.
taste processing at the molecular level. The 7-transmem-
brane-domain G protein-coupled receptors T1R1 and T1R3
Sour Taste
appear to form a heteromeric umami taste receptor (Li et al.,
2002; Nelson et al., 2002; Zhao et al., 2003). The T1R1/T1R3 Sour taste perception is generally agreed to be triggered
heteromer was first identified by expressing candidate taste when acidic substances stimulate the taste buds, causing de-
receptors in human embryonic kidney (HEK) cells and stim- polarization-induced Ca2 entry into taste receptor cells
ulating them with all 20 L-amino acids, as well as several D- (TRCs) (Richter et al., 2003). While slightly acidic foods are
amino acids that taste sweet to humans (Nelson et al., 2002). palatable to many animals (Kim et al., 2004), most mammals
Only cells coexpressing T1R1/T1R3 responded to L-amino reject strong sour stimuli, and it is thought that sour taste per-
acids, and their responses were potentiated by the umami ception may help prevent the consumption of spoiled foods
74 GARCIA-BAILO ET AL.

or serve as an indicator of fruit ripeness (DeSimone et al., 50% of the observed variance in sour stimulus perception
2001; Kim et al., 2004; Kinnamon and Margolskee, 1996; Lin- (Wise et al., 2007). This suggests a strong heritability com-
demann, 2001). Sources of sour tastants include inorganic ponent of sour taste sensitivity.
molecules such as hydrochloric acid and organic compounds Analyzing the genes for putative sour taste receptors, such
such as acetic, citric, lactic, or tartaric acid. Normally, these as the PKD2L1/PKD1L3 heteromer, could provide a start-
compounds are natural products of fermentation or basic ing point to explore inter-individual variation in this trait.
metabolic pathways such as the citric acid cycle. They can be Both PKD2L1 and PKD1L3 contain coding SNPs, and it is
found in most fruits and vegetables, as well as animal prod- possible that some of these polymorphisms may affect sour
ucts and man-made products such as wine (Roper, 2007). taste perception, but the potential relationship between poly-
The molecular machinery accounting for sour taste per- morphisms in these genes, sour taste perception, and subse-
ception in TRCs remained poorly understood for years. Pro- quent food choices remains to be explored.
posed mechanisms have included acid-sensing ion channels
found in rats (Ugawa et al., 1998), hyperpolarization-acti-
Salty Taste
vated cyclic nucleotide-gated channels (Stevens et al., 2001)
and pore domain potassium ion channels (Lin et al., 2004). The most abundant dietary source of salty taste is NaCl.
In the past few years, two transient receptor potential (TRP) A number of cations such as NH4, K, and Li also elicit
ion channels have gathered strong evidence as putative sour salty taste responses (DeSimone and Lyall, 2006; Roper,
taste receptors (Huang et al., 2006; Ishimaru et al., 2006; 2007), but their taste, rather than purely salty, appears to
LopezJimenez et al., 2006). These two ion channels, PKD2L1 be associated with bitterness, sourness, or astringency
and PKD1L3, belong to the polycystic kidney disease-like (Miyamoto et al., 2000; Roper, 2007). NaCl plays an essential
(PKDL) subfamily of TRPs, some of which act as nonselec- physiological role by maintaining electrolyte balance (Chan-
tive cation channels and are permeable to both Na and drashekar et al., 2006), as well as regulating blood pressure,
Ca2. Members of the PKD2L subfamily have a six trans- blood volume, and water homeostasis (Kim et al., 2004;
membrane domain, as well as a pore-forming domain. The Roper, 2007). NaCl and salty-tasting KCl contribute Na and
PDK1L proteins possess 11 transmembrane spanning do- K to the diet, which are ions that participate in important
mains, as well as large extracellular domains and short in- physiological processes such as nerve and muscle signalling,
tracellular carboxy domains (LopezJimenez et al., 2006). active transport across cell membranes, and maintaining cell
The coexpression of PKD2L1 and PKD1L3 was first ob- volume, pH, and cellular concentrations of other important
served in mouse taste cells, where it was noted that the two ions such as Ca2 (Sweeney and Klip, 2001).
molecules functioned as a heteromer and that they were ex- While the basis of salt taste perception has been studied
pressed in a different subset of taste cells than those for for years, its molecular mechanism remains unclear. In ro-
sweet, bitter, and umami (LopezJimenez et al., 2006). Subse- dents, epithelial sodium channels (ENaCs) located within
quently, PKD2L1 was located to the taste pore region of taste taste cell membranes in fungiform papillae may play an im-
buds, and transfected cells expressing both PKD2L1 and portant role in the perception of NaCl (Heck et al., 1984; Lin
PKD1L3 were shown to respond to acids, but not to other et al., 1999; Schiffman et al., 1983). Given the right concen-
tastants (Ishimaru et al., 2006). Finally, mice with genetically tration of Na in the oral cavity, Na flows passively through
ablated PKD2L1-expressing cells showed a nearly complete these ion channels, depolarizing the TRCs, and eliciting a salt
loss of sour taste perception, while their ability to detect taste response (Kim et al., 2004; Roper, 2007). ENaCs are sen-
other modalities remained unchanged (Huang et al., 2006). sitive to amiloride, a diuretic drug that inhibits Na trans-
These results highlight the role of PKD2L1 and PKD1L3 in port in various epithelial tissues (Schiffman et al., 1983).
sour taste perception. Nonetheless, it has been noted that, In humans, other mechanisms besides ENaCs may affect
while PKD2L1 is present in all taste cell types examined, NaCl perception. For example, the amiloride sensitivity of
PKD1L3 is absent from fungiform papillae and the palate NaCl perception appears to be specific to a minor sour com-
(Huang et al., 2006; Ishimaru et al., 2006). This observation ponent of this taste, rather than to the salty stimulus itself
could suggest that another, as yet unidentified molecule may (Ossebaard and Smith, 1995). Furthermore, salty taste per-
also play a role in sour taste reception. ception is inhibited in subjects that rinse with the oral anti-
Little is known about interindividual and interpopulation septic chlorhexidine, suggesting that salt-sensitive ion chan-
variation in sour taste perception, and how such variation nels in TRCs are not specifically sensitive to amiloride
may be linked to genetic variation. A European population (Breslin and Tharp, 2001).
was described as having a fairly narrow unimodal distribu- An amiloride-insensitive vanilloid receptor, Trpv1, has
tion of sensitivity to hydrochloric, citric, acetic, and picric been proposed to play a role in salty taste perception in ro-
acids (Blakeslee and Salmon, 1935). Existing twin studies dents (Lyall et al., 2004). This taste receptor has been hy-
provide equivocal information about the role of genetics in pothesized to respond to various cations, including Na, K,
sour taste perception (Kaplan et al., 1967; Wise et al., 2007). NH4, and Ca2 (DeSimone and Lyall, 2006). However, the
In monozygotic and fraternal twin pairs, there was little cor- importance of this protein has been questioned because
relation between absolute detection thresholds for hy- knockout mice lacking the receptor are nonetheless respon-
drochloric acid and degree of relatedness, suggesting that sive to salt taste (Ruiz et al., 2006). In absolute threshold de-
sour taste sensitivity was not a heritable trait (Kaplan et al., tection tests, Trpv1 knockout and wild-type mice detected
1967). However, by assessing the minimum concentrations NaCl, as well as KCl, at similar concentrations. The mice were
at which twins recognized the sour taste of citric and hy- also given preference tests of NaCl versus water, and of KCl
drochloric acid, rather than the absolute detection threshold, versus water. In both types of test, the Trpv1 knockout mice
it was found that additive genetic factors accounted for over actually preferred the salty tasting solution over water com-
GENETIC VARIATION IN TASTE 75

pared to wild-type animals. As these results suggest, the role tial role of NaCl in physiological processes, it is possible that
of Trpv1 in salty taste perception is unclear, but there seems evolutionary forces may have shaped the human ability to
to be agreement over the fact that salty taste is mediated by taste salt so as to be strongly influenced by environmental
ion flow, and both amiloride-sensitive and amiloride-insen- and dietary cues (Wise et al., 2007).
sitive sodium channels are involved in some capacity.
Genetic variation may affect salty taste perception in ro-
“Fat Taste”
dents, and different laboratory strains of mice and rats have
been observed to respond differently to NaCl and amiloride From an evolutionary perspective, “fat taste” perception
(Doolin and Gilbertson, 1996; Formaker and Hill, 1991; Ni- may have evolved to detect high energy foods and to select
nomiya et al., 1984a, 1984b). An SNP within the Scnn1a (so- foods containing fat soluble vitamins and essential fatty acids
dium channel, nonvoltage gated, type 1 ) gene, which en- (FAs) (Laugerette et al., 2007). Physiologically, fat detection
codes for the  ENaC subunit, has been linked to this in the cephalic phase may aid in preparing the digestive sys-
difference in response to NaCl and amiloride (Shigemura et tem for lipid metabolism. Indeed, in feeding trials an oral fat
al., 2008). This SNP (C1877T) results in an arginine to tryp- stimulus triggers a rapid increase in plasma triacylglycerol,
tophan substitution (R616W) in the  subunit protein. implicating incoming oral fat in the release of stored fat
Amiloride inhibits responses to NaCl more strongly in mice (Mattes, 2002). It is well known that in free choice scenarios
carrying a copy of the arginine-coding allele than in mice ho- there is a natural tendency toward fat consumption. Both rats
mozygous for the tryptophan-coding allele (Shigemura et al., and mice select a high fat diet over a low-fat diet (Hamilton,
2008). These results suggest that the R616W substitution in 1964) and humans may have a comparable inclination to-
the  ENaC subunit affects amiloride sensitivity of the ENaC ward dietary fat (Mela and Sacchetti, 1991). The mechanisms
channel, and may in turn affect salty taste responses in mice. guiding fat detection have traditionally been attributed to
In humans, variability in responses to salty stimuli has texture and olfaction. However, blocking the ability to sense
been examined for decades, but a direct genetic link to hu- these factors fails to abolish the recognition of dietary fat sug-
man salt taste perception has yet to be uncovered. A narrow, gesting another mechanism also contributes (Fukuwatari et
unimodal distribution of sensitivity to NaCl and KCl was al., 2003; Takeda et al., 2001). Our ability to detect trace
observed in a small sample of European subjects (Blakeslee amounts of FAs, including oleic, stearic, and linoleic FAs, on
and Salmon, 1935). However, an African population showed the tongue indicates that their oral detection may be another
a bimodal distribution of salty taste sensitivity (Odeigah, cue (Chale-Rush et al., 2007).
1994). While this may indicate a heritability component, en- Some of the early evidence for a “fat taste” receptor came
vironmental exposure to NaCl probably plays an important from observations that FAs inhibit delayed rectifying K
role in observed salty taste perception variability (Crystal channels on TRC, prolonging cell depolarization (Gilbertson
and Bernstein, 1995; Pittman and Contreras, 2002; Stein et et al., 1997). The response appeared to be specific to unsat-
al., 1996; Wise et al., 2007). Furthermore, as evidenced by urated long-chain fatty acids (LCFAs) and was generally
twin studies, salty taste perception appears to be determined limited to stimuli applied extracellularly, consistent with an
more by environment than by heritability components apical taste receptor. One good candidate for a “fat taste” re-
(Beauchamp, et al., 1985; Wise et al., 2007). Given the essen- ceptor is CD36, which Laugerette et al. (2005) have recently

FIG. 2. With the discovery of the putative taste receptors for the different taste modalities (bitter, sweet, umami, sour,
salty, and fat), we can begin to examine genetic variation leading to differences in taste perception. The identification of ge-
netic polymorphisms associated with differences in taste perception will allow for functionality studies to determine phe-
notypic outcomes and assess their impact on food selection. Food selection results from the balance between taste percep-
tion and other factors, including nutritional status, physiology, environment, and sociocultural factors.
76 GARCIA-BAILO ET AL.

proposed to be an oral lipid sensor. CD36 is an integral mem- Discussion and Conclusion
brane protein that has a high affinity for LCFA (Baillie et al.,
The characterization of taste receptor genes, together with
1996) and is generally known for its role in facilitating free
the development of genomic technologies, has generated
fatty acid (FFA) transport across the cell membrane (Harmon
new avenues of research on the physiology of taste in hu-
and Abumrad, 1993). However, some controversy surrounds
mans. Different approaches can be used to determine how
this role (Hamilton, 2007). CD36 was isolated on the apical
variations in taste receptor genes influence taste perception,
surface of taste bud cells the same year as the K channel
food choices, and dietary intake behaviors. Understanding
TRC studies (Fukuwatari et al., 1997). The protein spans the
this relationship requires distinguishing actual dietary in-
cell membrane creating a large extracellular hydrophobic
take from food preferences. Using different methods of di-
loop, likely the portion that interacts with FAs, and two short
etary assessment such as food frequency questionnaires,
cytoplasmic tails (Abumrad et al., 1993). Although this pro-
food diaries, and food preference checklists can help estab-
tein structure expressed on taste buds supports an apical
lish the validity and reliability of genetic markers of dietary
taste receptor, the precise mechanism by which the protein
intake. Threshold studies that incorporate genotype may
senses FAs is not clear. The sensing ability appears to be spe-
help characterize the functionality of specific polymor-
cific to longer chain FAs, but both saturated and unsaturated
phisms as either having a low detection threshold or induc-
LCFA can cause depolarizing increases in intracellular cal-
ing a supertasting phenotype for a particular taste modal-
cium in CD36-positive cells. Moreover, a CD36-specific in-
ity. Because other biological factors are involved in food
hibitor, sulfo-N-succinimidyl oleic acid ester, attenuates this
selection, future studies should examine how taste percep-
response (Gaillard et al., 2008). In addition to being ex-
tion interacts with energy homeostatic pathways and food
pressed on the tongue, CD36 is also expressed in the stom-
reward circuits for a comprehensive understanding of in-
ach and intestine and on the surface of macrophages,
gestive behavior. It is also important to consider how non-
adipocytes, muscle cells, endothelial cells, and platelets.
genetic aspects such as environment and sociocultural fac-
Direct evidence demonstrating that CD36 may act as a pu-
tors influence food selection. The contributions from each
tative “fat taste” receptor comes from animal studies using
of these variables determine overall ingestive behavior, with
knockout mice (Laugerette et al., 2005). Unlike wild-type mice
the potential for nongenetic factors to override genetic pre-
that normally choose a FFA containing diet over control, mice
disposition (Fig. 2).
with a targeted deletion of the CD36 gene lose the ability to
Genetic and environmental factors contribute to nutri-
distinguish between the two (Laugerette et al., 2005). These
tional and overall health status, which are both growing pub-
observations appear to be specific for fat, as opposed to a gen-
lic health concerns. Understanding food intake behaviors
eral loss of taste, as indicated by the maintenance of a pref-
from the perspective of taste perception is important because
erence for sweet and aversion for bitter (Laugerette et al.,
it lies at the interface between the foods an individual is ex-
2005). Although dietary fat is predominantly in the form of
posed to and the biological predisposition to prefer certain
triglycerides, lingual lipase may yield sufficient FFA to act as
foods in one’s environment. Examining variations in taste re-
a chemical cue. Kawai and Fushiki (2003) showed that inhi-
ceptors will help establish the association between certain
bition of lingual lipase reduced the intake of triglyceride, but
food intake behaviors and risk of chronic disease. Polymor-
failed to reduce the intake of FFA. Indeed, lipase secretion is
phisms in taste receptors might be useful as surrogate mark-
constant and yields FFAs quickly enough to enable them to
ers of dietary exposure in gene–disease association studies
be detected by a fat sensor on the surface of the tongue (Kawai
where information on dietary habits is not available. In ad-
and Fushiki, 2003). A more recent study confirmed that com-
dition, differences in genetic variation in taste receptors be-
pared to lipid naïve CD36-null mice, wild-type mice are more
tween ethnic groups may contribute to differences in dietary
likely to choose a FA solution over gum vehicle (0.3% xan-
intake patterns. Understanding these differences in genetic
than gum), indicating that CD36 is required to distinguish
variation could help explain the interethnic disparities in risk
these texturally comparable choices (Sclafani et al., 2007) .
for chronic disease and lead to the development of appro-
These CD36-null mice also showed a weaker preference for
priate preventative public health measures.
triglyceride over Emplex vehicle (0.15% sodium stearoyl
lactylate) than their wild-type littermates (Sclafani et al.,
Acknowledgments
2007). However, at increasing concentrations of both FA and
triglyceride CD36-null mice exhibited increased fat prefer- This work was supported by a grant from the Advanced
ences, although their total fat intake continued to be less than Foods & Materials Network (AFMNet). A. El-Sohemy holds
that of their wild-type littermates. The authors suggest that a Canada Research Chair in Nutrigenomics.
postoral conditioning effects may in part be responsible for
this “rescued” phenotype (Sclafani et al., 2007). Author Disclosure Statement
The role of CD36 in mediating fat intake in humans is not
yet known. A number of sequence variations in the human The authors declare that no competing financial interests
CD36 gene, located on chromosome 7q11.2, have been iden- exist.
tified (Fernandez-Ruiz et al., 1993). If genetic variation in
CD36 affects our ability to sense or taste FFA it may result References
in variation in preferences for fatty foods. Thus, examining Abumrad, N.A., El-Maghrabi, M.R., Amri, E.Z., Lopez, E., and
the relationship between inherited variations in CD36 with Grimaldi, P.A. (1993). Cloning of a rat adipocyte membrane
fat consumption and oral chemosensory response to fat may protein implicated in binding or transport of long-chain fatty
help identify individuals predisposed to prefer foods higher acids that is induced during preadipocyte differentiation. Ho-
in dietary fat. mology with human CD36. J Biol Chem 268, 17665–17668.
GENETIC VARIATION IN TASTE 77

Adler, E., Hoon, M.A., Mueller, K.L., Chandrashekar, J. Ryba, Chang, W.I., Chung, J.W., Kim, Y.K., Chung, S.C., and Kho, H.S.
N.J., and Zuker, C.S. (2000). A novel family of mammalian (2006). The relationship between phenylthiocarbamide (PTC)
taste receptors. Cell 100, 693–702. and 6–n-propylthiouracil (PROP) taster status and taste
Anliker, J.A., Bartoshuk, L.M., Ferris, A.M., and Hooks, L.D. thresholds for sucrose and quinine. Arch Oral Biol 51, 427–432.
(1991). Children’s food preferences and genetic sensitivity to Chaudhari, N., and Roper, S.D. (1998). Molecular and physio-
the bitter tastes of 6–n-propylthiouracil (PROP). Am J Clin logical evidence for glutamate (umami) taste transduction via
Nutr 54, 316–320. a G protein-coupled receptor. Ann N Y Acad Sci 855, 398–406.
Bachmanov, A.A., and Beauchamp, G.K. (2007). Taste receptor Chaudhari, N., Yang, H., Lamp, C., Delay, E., Cartford, C., Than,
genes. Annu Rev Nutr 27, 389–414. T., et al. (1996). The taste of monosodium glutamate: mem-
Bachmanov, A.A., Li, X., Reed, D.R., Ohmen, J.D., Li, S., Chen, brane receptors in taste buds. J Neurosci 16, 3817–3826.
Z., et al. (2001). Positional cloning of the mouse saccharin pref- Chaudhari, N., Landin, A.M., and Roper, S.D. (2000). A metabo-
erence (Sac) locus. Chem Senses 26, 925–933. tropic glutamate receptor variant functions as a taste recep-
Bachmanov, A.A., Reed, D.R., Li, X., and Beauchamp, G.K. tor. Nat Neurosci 3, 113–119.
(2002). Genetics of sweet taste preferences. Pure Appl Chem Crystal, S.R., and Bernstein, I.L. (1995). Morning sickness: im-
74, 1135–1140. pact on offspring salt preference. Appetite 25, 231–240.
Baillie, A.G., Coburn, C.T., and Abumrad, N.A. (1996). Re- Delay, E.R., Hernandez, N.P., Bromley, K., and Margolskee,
versible binding of long-chain fatty acids to purified FAT, the R.F. (2006). Sucrose and monosodium glutamate taste thresh-
adipose CD36 homolog. J Membr Biol 153, 75–81. olds and discrimination ability of T1R3 knockout mice. Chem
Barnicot, N.A., Harris, H., and Kalmus, H. (1951). Taste thresh- Senses 31, 351–357.
olds of further eighteen compounds and their correlation with Desimone, J.A., and Lyall, V. (2006). Taste receptors in the gas-
P.T.C thresholds. Ann Eugen 16, 119–128. trointestinal tract III. Salty and sour taste: sensing of sodium
Bartoshuk, L.M. (2000). Comparing sensory experiences across and protons by the tongue. Am J Physiol Gastrointest Liver
individuals: recent psychophysical advances illuminate ge- Physiol 291, G1005–G1010.
netic variation in taste perception. Chem Senses 25, 447–460. Desimone, J.A., Lyall, V., Heck, G.L., and Feldman, G.M. (2001).
Bartoshuk, L.M., Duffy, V.B., and Miller, I.J. (1994). PTC/PROP Acid detection by taste receptor cells. Respir Physiol 129,
tasting: anatomy, psychophysics, and sex effects. Physiol Be- 231–245.
hav 56, 1165–1171. Doolin, R.E., and Gilbertson, T.A. (1996). Distribution and char-
Bartoshuk, L.M., Duffy, V.B., Lucchina, L.A., Prutkin, J., and acterization of functional amiloride-sensitive sodium channels
Fast, K. (1998). PROP (6–n-propylthiouracil) supertasters and in rat tongue. J Gen Physiol 107, 545–554.
the saltiness of NaCl. Ann N Y Acad Sci 855, 793–796. Drayna, D. (2005). Human taste genetics. Annu Rev Genomics
Basson, M.D., Bartoshuk, L.M., Dichello, S.Z., Panzini, L., Weif- Hum Genet 6, 217–235.
fenbach, J.M., and Duffey, V.B. (2005). Association between Drewnowski, A., and Rock, C.L. (1995). The influence of ge-
6-n-propylthiouracil (PROP) bitterness and colonic neoplasms. netic taste markers on food acceptance. Am J Clin Nutr 62,
Dig Dis Sci 50, 483–489. 506–511.
Beauchamp, G., Bertino, M., and Engelman, K. (1985). Sensory Drewnowski, A., Henderson, S.A., Shore, A.B., and Barratt-For-
basis of human salt consumption. In NIH Workshop on Nutri- nell, A. (1998). Sensory responses to 6-n-propylthiouracil
tion and Hypertension. M. Horan, M. Blaustein, J. Dunbar, W. (PROP) or sucrose solutions and food preferences in young
Kachadorian, N. Kaplan, and A. Simopoulos, Eds. (Biomed. women. Ann N Y Acad Sci 855, 797–801.
Info. Corp., New York). Drewnowski, A., Henderson, S.A., Levine, A., and Hann, C.S.
Blakeslee, A.F. (1932). Genetics of sensory thresholds: taste for (1999). Taste and food preferences as predictors of dietary
phenyl thio carbamide. Proc Natl Acad Sci USA 18, 120–130. practices in young women. Public Health Nutr 2, 513–519.
Blakeslee, A.F., and Salmon, T.N. (1935). Genetics of sensory Drewnowski, A., Henderson, S.A., and Barratt-Fornell, A. (2001).
thresholds: individual taste reactions for different substances. Genetic taste markers and food preferences. Drug Metab Dis-
Proc Natl Acad Sci USA 21, 84–90. pos 29, 535–538.
Boughter, J.D., Jr., and Bachmanov, A.A. (2007). Behavioral ge- Duffy, V.B. (2004). Associations between oral sensation, dietary
netics and taste. BMC Neurosci 8(Suppl 3), S3. behaviors and risk of cardiovascular disease (CVD). Appetite
Breslin, P.A., and Tharp, C.D. (2001). Reduction of saltiness and 43, 5–9.
bitterness after a chlorhexidine rinse. Chem Senses 26, 105–116. Duffy, V.B., and Bartoshuk, L.M. (2000). Food acceptance and
Bufe, B., Hofmann, T., Krautwurst, D., Raguse, J.D., and Mey- genetic variation in taste. J Am Diet Assoc 100, 647–655.
erhof, W. (2002). The human TAS2R16 receptor mediates bit- Duffy, V., Davidson, A.C., Kidd, J.R., Kidd, K.K., Speed, W.C.,
ter taste in response to beta-glucopyranosides. Nat Genet 32, Pakstis, A.J., et al. (2004). Bitter receptor gene (TAS2R38), 6-
397–401. n-propylthiouracil (PROP) bitterness and alcohol intake. Al-
Bufe, B., Breslin, P.A., Kuhn, C., Reed, D.R., Tharp, C.D., Slack, cohol Clin Exp Res 28, 1629–1637.
J.P., et al. (2005). The molecular basis of individual differences El-Sohemy, A., Stewart, L., Khataan, N., Fontaine-Bisson, B.,
in phenylthiocarbamide and propylthiouracil bitterness per- Kwong, P., Ozsungur, S., et al. (2007). Nutrigenomics of
ception. Curr Biol 15, 322–327. taste—impact on food preferences and food production. Fo-
Chale-Rush, A., Burgess, J.R., and Mattes, R.D. (2007). Evidence rum Nutr 60, 176–182.
for human orosensory (taste?) sensitivity to free fatty acids. Eny, K.M., Wolever, T.M., Fontaine-Bisson, B., and El-Sohemy,
Chem Senses 32, 423–431. A. (2008). Genetic variant in the glucose transporter type 2
Chandrashekar, J., Mueller, K.L., Hoon, M.A., Adler, E., Feng, (GLUT2) is associated with higher intakes of sugars in two
L., Guo, W., et al. (2000). T2Rs function as bitter taste recep- distinct populations. Physiol Genomics 33, 355–360.
tors. Cell 100, 703–711. Fernandez-Ruiz, E., Armesilla, A.L., Sanchez-Madrid, F., and
Chandrashekar, J., Hoon, M.A., Ryba, N.J., and Zuker, C.S. Vega, M.A. (1993). Gene encoding the collagen type I and
(2006). The receptors and cells for mammalian taste. Nature thrombospondin receptor CD36 is located on chromosome
444, 288–294. 7q11.2. Genomics 17, 759–761.
78 GARCIA-BAILO ET AL.

Formaker, B.K., and Hill, D.L. (1991). Lack of amiloride sensi- Huang, A.L., Chen, X., Hoon, M.A., Chandrashekar, J., Guo, W.,
tivity in SHR and WKY glossopharyngeal taste responses to Trankner, D., et al. (2006). The cells and logic for mammalian
NaCl. Physiol Behav 50, 765–769. sour taste detection. Nature 442, 934–938.
Fox, A.L. (1931). Six in ten “tasteblind” to bitter chemical. Sci Ikeda, K. (2002). New seasonings. Chem Senses 27, 847–849.
News l 9, 249. Ishimaru, Y., Inada, H., Kubota, M., Zhuang, H., Tominaga, M.,
Fukuwatari, T., Kawada, T., Tsuruta, M., Hiraoka, T., Iwanaga, and Matsunami, H. (2006). Transient receptor potential fam-
T., Sugimoto, E., et al. (1997). Expression of the putative mem- ily members PKD1L3 and PKD2L1 form a candidate sour taste
brane fatty acid transporter (FAT) in taste buds of the cir- receptor. Proc Natl Acad Sci USA 103, 12569–12574.
cumvallate papillae in rats. FEBS Lett 414, 461–464. Kaminski, L.C., Henderson, S.A., and Drewnowski, A. (2000).
Fukuwatari, T., Shibata, K., Iguchi, K., Saeki, T., Iwata, A., Tani, Young women’s food preferences and taste responsiveness to
K., et al. (2003). Role of gustation in the recognition of oleate 6-n-propylthiouracil (PROP). Physiol Behav 68, 691–697.
and triolein in anosmic rats. Physiol Behav 78, 579–583. Kaplan, A.R., Fischer, R., Karras, A., Griffin, F., Powell, W.,
Fuller, J.L. (1974). Single-locus control of saccharin preference in Marsters, R.W., et al. (1967). Taste thresholds in twins and sib-
mice. J Hered 65, 33–36. lings. Acta Genet Med Gemellol (Roma) 16, 229–243.
Gaillard, D., Laugerette, F., Darcel, N., El-Yassimi, A., Passilly- Kataoka, S., Yang, R., Ishimaru, Y., Matsunami, H., Sevigny, J.,
Degrace, P., Hichami, A., et al. (2008). The gustatory path- Kinnamon, J.C., et al. (2008). The candidate sour taste recep-
way is involved in CD36–mediated orosensory perception tor, PKD2L1, is expressed by type III taste cells in the mouse.
of long-chain fatty acids in the mouse. FASEB J 22, 1458– Chem Senses, 33, 243–254.
1468. Kawai, T., and Fushiki, T. (2003). Importance of lipolysis in oral
Gilbertson, T.A., Fontenot, D.T., Liu, L., Zhang, H., and Mon- cavity for orosensory detection of fat. Am J Physiol Regul In-
roe, W.T. (1997). Fatty acid modulation of K channels in taste tegr Comp Physiol 285, R447–R454.
receptor cells: gustatory cues for dietary fat. Am J Physiol 272, Keller, K.L., Steinmann, L., Nurse, R.J., and Tepper, B.J. (2002).
C1203–C1210. Genetic taste sensitivity to 6-n-propylthiouracil influences
Glanz, K., Basil, M., Maibach, E., Goldberg, J., and Snyder, D. food preference and reported intake in preschool children.
(1998). Why Americans eat what they do: taste, nutrition, cost, Appetite 38, 3–12.
convenience, and weight control concerns as influences on Keskitalo, K., Tuorila, H., Spector, T.D., Cherkas, L.F., Knaapila,
food consumption. J Am Diet Assoc 98, 1118–1126. A., Silventoinen, K., et al. (2007a). Same genetic components
Guo, S.W., and Reed, D.R. (2001). The genetics of phenylthio- underlie different measures of sweet taste preference. Am J
carbamide perception. Ann Hum Biol 28, 111–142. Clin Nutr 86, 1663–1669.
Hall, M.J., Bartoshuk, L.M., Cain, W.S., and Stevens, J.C. (1975). Keskitalo, K., Knaapila, A., Kallela, M., Palotie, A., Wessman,
PTC taste blindness and the taste of caffeine. Nature 253, M., Sammalisto, S., et al. (2007b). Sweet taste preferences are
442–443. partly genetically determined: identification of a trait locus on
Hamilton, C.L. (1964). Rat’s preference for high fat diets. J Comp chromosome 16. Am J Clin Nutr 86, 55–63.
Physiol Psychol 58, 459–460. Kim, U.K., and Drayna, D. (2004). Genetics of individual differ-
Hamilton, J.A. (2007). New insights into the roles of proteins and ences in bitter taste perception: lessons from the PTC gene.
lipids in membrane transport of fatty acids. Prostaglandins Clin Genet 67, 275–280.
Leukot Essent Fatty Acids 77, 355–361. Kim, U.K., Jorgenson, E., Coon, H., Leppert, M., Risch, N., and
Harmon, C.M., and Abumrad, N.A. (1993). Binding of sulfos- Drayna, D. (2003). Positional cloning of the human quantita-
uccinimidyl fatty acids to adipocyte membrane proteins: iso- tive trait locus underlying taste sensitivity to phenylthiocar-
lation and amino-terminal sequence of an 88-kD protein im- bamide. Science 299, 1221–1225.
plicated in transport of long-chain fatty acids. J Membr Biol Kim, U.K., Breslin, P.A., Reed, D., and Drayna, D. (2004). Ge-
133, 43–49. netics of human taste perception. J Dent Res 83, 448–453.
Hayes, J.E., Bartoshuk, L.M., Kidd, J.R., and Duffy, V.B. (2008). Kim, U.K., Wooding, S., Riaz, N., Jorde, LB., and Drayna, D.
Supertasting and PROP bitterness depends on more than the (2006). Variation in the human TAS1R taste receptor genes.
TAS2R38 gene. Chem Senses 33, 255–265. Chem Senses 31, 599–611.
Heck, G.L., Mierson, S., and Desimone, J.A. (1984). Salt taste Kinnamon, S.C., and Margolskee, R.F. (1996). Mechanisms of
transduction occurs through an amiloride-sensitive sodium taste transduction. Curr Opin Neurobiol 6, 506–513.
transport pathway. Science 223, 403–405. Kitagawa, M., Kusakabe, Y., Miura, H., Ninomiya, Y., and Hino,
Henkin, R.I., and Shallenberger, R.S. (1970). Aglycogeusia: the A. (2001). Molecular genetic identification of a candidate re-
inability to recognize sweetness and its possible molecular ba- ceptor gene for sweet taste. Biochem Biophys Res Commun
sis. Nature 227, 965–966. 283, 236–242.
Hinrichs, A.L., Wang, J.C., Bufe, B., Kwon, J.M., Budde, J., Allen, Knight, I. (1994). The development and applications of sucralose,
R., et al. (2006). Functional variant in a bitter-taste receptor a new high-intensity sweetener. Can J Physiol Pharmacol 72,
(hTAS2R16) influences risk of alcohol dependence. Am J Hum 435–439.
Genet 78, 103–111. Kuhn, C., Bufe, B., Winnig, M., Hofmann, T., Frank, O., Behrens,
Hladik, C.M., Pasquet, P., and Simmen, B. (2002). New per- M., et al. (2004). Bitter taste receptors for saccharin and ace-
spectives on taste and primate evolution: the dichotomy in sulfame K. J Neurosci 24, 10260–10265.
gustatory coding for perception of beneficent versus noxious Kurihara, K., and Kashiwayanagi, M. (2000). Physiological stud-
substances as supported by correlations among human ies on umami taste. J Nutr 130, 931S-934S.
thresholds. Am J Phys Anthropol 117, 342–348. Laugerette, F., Passilly-Degrace, P., Patris, B., Niot, I., Febbraio,
Hoon, M.A., Adkler, E., Lindemeier, J., Battery, J.F., Ryba, N.Y., M., Montmayeur, J.P., et al. (2005). CD36 involvement in
and Zuker, C.S. (1999). Putative mammalian taste receptors: orosensory detection of dietary lipids, spontaneous fat
a class of taste-specific GPCRs with distinct topographic se- preference, and digestive secretions. J Clin Invest 115, 3177–
lectivity. Cell 96, 541–551. 3184.
GENETIC VARIATION IN TASTE 79

Laugerette, F., Gaillard, D., Passilly-Degrace, P., Niot, I., and Miyamoto, T., Fujiyama, R., Okada, Y., and Sato, T. (2000). Acid
Besnard, P. (2007). Do we taste fat? Biochimie 89, 265–269. and salt responses in mouse taste cells. Prog Neurobiol 62,
Leterme, A., Brun, L., Dittmar, A., and Robin, O. (2008). Auto- 135–157.
nomic nervous system responses to sweet taste: Evidence for Montmayeur, J.P., Liberles, S.D., Matsunami, H., and Buck, L.B.
habituation rather than pleasure. Physiol Behav. in press. (2001). A candidate taste receptor gene near a sweet taste lo-
Li, X., Staszewski, L., Xu, H., Durick, K. Zoller, M., and Adler, cus. Nat Neurosci 4, 492–498.
E. (2002). Human receptors for sweet and umami taste. Proc Navarro-Allende, A., Khataan, N., and El-Sohemy, A. (in press).
Natl Acad Sci USA 99, 4692–4696. Impact of genetic and environmental determinants of taste
Li., X., Li, W., Wang, H., Cao, J., Maehashi, K., Huang, L., et al. with food preferences in older adults. J Nutr Elderly.
(2005). Pseudogenization of a sweet-receptor gene accounts Nelson, G., Hoon, M.A., Chandrashekar, J., Zhang, Y., Ryga, N.J.,
for cats’ indifference toward sugar. PLoS Genet 1, 27–35. Zuker, C.S. (2001). Mammalian sweet taste receptors. Cell 106,
Liao, J., and Schultz, P.G. (2003). Three sweet receptor genes are 381–390.
clustered in human chromosome 1. Mamm Genome 14, Nelson, G., Chandrashekar, J., Hoon, M.A., Feng, L., Zhao, G.,
291–301. Ryba, N.J., et al. (2002). An amino-acid taste receptor. Nature
Lin, W., Finger, T.E., Rossier, B.C., and Kinnamon, S.C. (1999). 416, 199–202.
Epithelial Na channel subunits in rat taste cells: localization Nie, Y., Vigues, S., Hobbs, J.R., Conn, G.L., and Munger, S.D.
and regulation by aldosterone. J Comp Neurol 405, 406–420. (2005). Distinct contributions of T1R2 and T1R3 taste receptor
Lin, W., Burks, C.A., Hansen, D.R., Kinnamon, S.C., and Gilbert- subunits to the detection of sweet stimuli. Curr Biol 15,
son, T.A. (2004). Taste receptor cells express pH-sensitive leak 1948–1952.
K channels. J Neurophysiol 92, 2909–2919. Ninomiya, Y., Higashi, T., Katsukawa, H., Mizukoshi, T., and
Lindemann, B. (2001). Receptors and transduction in taste. Na- Funakoshi, M. (1984a). Qualitative discrimination of gustatory
ture 413, 219–225. stimuli in three different strains of mice. Brain Res 322, 83–92.
Lindemann, B., Ogiwara, Y., and Ninomiya, Y. (2002). The dis- Ninomiya, Y., Mizukoshi, T., Higashi, T., Katsukawa, H., and
covery of umami. Chem Senses 27, 843–844. Funakoshi, M. (1984b). Gustatory neural responses in three
LopezJimenez, N.D., Cavenagh, M.M., Sainz, E., Cruz-Ithier, different strains of mice. Brain Res 302, 305–314.
M.A., Battey, J.F., and Sullivan, S.L. (2006). Two members of Odeigah, P.G. (1994). Smell acuity for acetone and its relation-
the TRPP family of ion channels, Pkd1l3 and Pkd2l1, are co- ship to taste ability to phenylthiocarbamide in a Nigerian pop-
expressed in a subset of taste receptor cells. J Neurochem 98, ulation. East Afr Med J 71, 462–466.
68–77. Ossebaard, C.A., and Smith, D.V. (1995). Effect of amiloride on
Lugaz, O., Pillias, A.M., and Faurion, A. (2002). A new specific the taste of NaCl, Na-gluconate and KCl in humans: implica-
ageusia: some humans cannot taste L-glutamate. Chem Senses tions for Na receptor mechanisms. Chem Senses 20, 37–46.
27, 105–115. Parker, K.J. (1978). Alternatives to sugar. The search for an ideal
Lyall, V., Heck, G.L., Vinnikova, A.K., Ghosh, S., Phan, T.H., non-nutritive sweetener is almost a century old. Nature 271,
Alam, R.I., et al. (2004). The mammalian amiloride-insensitive 493–495.
non-specific salt taste receptor is a vanilloid receptor-1 vari- Pickering, G., Haverstock, G., and Dibattista, D. (2006). Evidence
ant. J Physiol 558, 147–159. that sensitivity to 6–n-propylthiouracil (PROP) affects per-
Mace, O.J., Affleck, J., Patel, N., and Kellett, G.I. (2007). Sweet ception of retro-nasal aroma intensity. J Food Agric Environ
taste receptors in rat small intestine stimulate glucose ab- 4, 15–22.
sorption through apical GLUT2. J Physiol 582, 379–392. Pittman, D.W., and Contreras, R.J. (2002). Rearing on basal or
Margolskee, R.F. (2002). Molecular mechanisms of bitter and high dietary NaCl modifies chorda tympani nerve responses
sweet taste transduction. J Biol Chem 277, 1–4. in rats. Physiol Behav 77, 277–289.
Margolskee, R.F., Dyer, J., Kokrashvili, Z., Salmon, K.S., Ilegems, Prescott, J., and Swain-Campbell, N. (2000). Responses to re-
E., Daly, K., et al. (2007). T1R3 and gustducin in gut sense sug- peated oral irritation by capsaicin, cinnamaldehyde and eth-
ars to regulate expression of Na-glucose cotransporter 1. anol in PROP tasters and non-tasters. Chem Senses 25,
Proc Natl Acad Sci USA 104, 15075–15080. 239–246.
Matsunami, H., Montemayeur, J.P., and Buck, L.B. (2000). A fam- Prescott, J., Soo, J., Campbell, H., and Roberts, C. (2004). Re-
ily of candidate taste receptors in human and mouse. Nature sponses of PROP taster groups to variations in sensory qual-
404, 601–604. ities within foods and beverages. Physiol Behav 82, 459–469.
Mattes, R.D. (2002). Oral fat exposure increases the first phase Richter, T.A., Caicedo, A., and Roper, S.D. (2003). Sour taste
triacylglycerol concentration due to release of stored lipid in stimuli evoke Ca2 and pH responses in mouse taste cells. J
humans. J Nutr 132, 3656–3662. Physiol 547, 475–483.
Max, M., Shanker, Y.G., Huang, L., Rong, M., Liu, Z., Campagne, Roper, S.D. (2007). Signal transduction and information pro-
F., et al. (2001). Tas1r3, encoding a new candidate taste re- cessing in mammalian taste buds. Pflugers Arch 454, 759–776.
ceptor, is allelic to the sweet responsiveness locus Sac. Nat Ruiz, C., Gutknecht, S., Delay, E., and Kinnamon, S. (2006). De-
Genet 28, 58–63. tection of NaCl and KCl in TRPV1 knockout mice. Chem
Mazur, R. (1984). Discovery of Aspartame. In Aspartame: Physi- Senses 31, 813–820.
ology and Biochemistry. L. Stegink and L. Filer, eds. (Marcel Sacerdote, C., Guarrera, S., Smith, G.D., Grioni, S., Krogh, V.,
Dekker, New York). Masala, G., et al. (2007). Lactase persistence and bitter taste
Mela, D.J., and Sacchetti, D.A. (1991). Sensory preferences for response: instrumental variables and mendelian randomiza-
fats: relationships with diet and body composition. Am J Clin tion in epidemiologic studies of dietary factors and cancer risk.
Nutr 53, 908–915. Am J Epidemiol 166, 576–581.
Mennella, J.A., Pepino, M.Y., and Reed, D.R. (2005). Genetic and Sainz, E., Korley, J.N., Battey, J.F., and Sullivan, S.L. (2001). Iden-
environmental determinants of bitter perception and sweet tification of a novel member of the T1R family of putative taste
preferences. Paediatrics 115, 216–222. receptors. J Neurochem 77, 896–903.
80 GARCIA-BAILO ET AL.

Sainz, E., Cavenagh, M.M., Lopezjimenez, N.D., Gutierrez, J.C., Timpson, N.J., Christensen, M., Lawlor, D.A., Gaunt, T.R., Day,
Battey, J.F., Northup, J.K., et al. (2007). The G-protein coupling L.N., and Davey Smith, G. (2005). TAS2R38 (phenylthiocar-
properties of the human sweet and amino acid taste receptors. bamide) haplotypes, coronary heart disease traits, and eating
Dev Neurobiol 67, 948–959. behavior in the British Women’s Heart and Health Study. Am
Schiffman, S.S., Lockhead, E., and Maes, F.W. (1983). Amiloride J Clin Nutr 81, 1005–1011.
reduces the taste intensity of Na and Li salts and sweet- Tordoff, M.G. (1988). How do non-nutritive sweeteners increase
eners. Proc Natl Acad Sci USA 80, 6136–6140. food intake? Appetite 11(Suppl 1), 5–11.
Sclafani, A., Ackroff, K., and Abumrad, N.A. (2007). CD36 gene Turnbull, B., and Matisoo-Smith, E. (2002). Taste sensitivity to
deletion reduces fat preference and intake but not post-oral 6–n-propylthiouracil (PROP) predicts acceptance of bitter-
fat conditioning in mice. Am J Physiol Regul Integr Comp tasting spinach in 3–6 year old children. Am J Clin Nutr 76,
Physiol 293, R1823–R1832. 1101–1105.
Shiffman, D., Ellis, S.G., Rowland, C.M., Malloy, M.J., Luke, Ugawa, S., Minami, Y., Guo, W., Saishin, Y., Takatsuji, K., Yam-
M.M., Iakoubova, O. A., et al. (2005). Identification of four mamoto, T., et al. (1998). Receptor that leaves a sour taste in
gene variants associated with myocardial infarction. Am J the mouth. Nature 395, 555–556.
Hum Genet 77, 596–605. Wang, J.C., Hinrichs, A.L., Bertelsen, S., Stock, H., Budde, J.P.,
Shiffman, D., O’Meara, E.S., Bare, L.A., Rowland, C.M., Louie, dick, D.M., et al. (2007). Functional variants in TAS2R38 and
J.Z., Arellano, A.R., et al. (2008). Association of gene variants TAS2R16 influence alcohol consumption in high-risk families
with incident myocardial infarction in the Cardiovascular of African-American origin. Alcohol Clin Exp Res 31, 209–
Health Study. Arterioscler Thromb Vasc Biol 28, 173–179. 215.
Shigemura, N., Ohkuri, T., Sadamitsu, C., Yasumatsu, K., Whissell-Buechy, D. (1990). Effects of age and sex on taste sen-
Yoshida, R., Beauchamp, G.K., et al. (2008). Amiloride-sensi- sitivity to phenylthiocarbamide (PTC) in the Berkeley Guid-
tive NaCl taste responses are associated with genetic varia- ance sample. Chem Senses 15, 39–57.
tion of ENaC alpha-subunit in mice. Am J Physiol Regul In- Wise, P.M., Hansen, J.L., Reed, D.R., and Breslin, P.A. (2007).
tegr Comp Physiol 294, R66–R75. Twin study of the heritability of recognition thresholds for
Stein, L.J., Cowart, B.J., Epstein, A.N., Pilot, L.J., Laskin, C.R., sour and salty taste. Chem Senses 32, 749–754.
and Beauchamp, G.K. (1996). Increased liking for salty foods Xu, H., Staszewski, L., Tang, H., Adler, E., Zoller, M., and Li, X.
in adolescents exposed during infancy to a chloride-deficient (2004). Different functional roles of T1R subunits in the het-
feeding formula. Appetite 27, 65–77. eromeric taste receptors. Proc Natl Acad Sci USA 101,
Stevens, D.R., Seifert, R., Bufe, B., Muller, F., Kremmer, E., Gauss, 14258–14263.
R., et al. (2001). Hyperpolarization-activated channels HCN1 and Zafra, M.A., Molina, F., and Puerto, A. (2006). The neural/
HCN4 mediate responses to sour stimuli. Nature 413, 631–635. cephalic phase reflexes in the physiology of nutrition. Neu-
Sweeney, G., and Klip, A. (2001). Mechanisms and consequences rosci Biobehav Rev 30, 1032–1044.
of Na,K-pump regulation by insulin and leptin. Cell Mol Zhao, G.Q., Zhang, Y., Hoon, M.A., Chandrashekar, J., Erlen-
Biol (Noisy-le-grand) 47, 363–372. bach, I., and Lyba, N.J. (2003). The receptors for mammalian
Swithers, S.E., and Davidson, T.L. (2008). A role for sweet taste: sweet and umami taste. Cell 115, 255–266.
calorie predictive relations in energy regulation by rats. Be-
hav Neurosci 122, 161–173.
Takeda, M., Sawano, S., Imaizumi, M., and Fushiki, T. (2001). Address reprint requests to:
Preference for corn oil in olfactory-blocked mice in the con- Dr. Ahmed El-Sohemy
ditioned place preference test and the two-bottle choice test.
Department of Nutritional Sciences
Life Sci 69, 847–854.
Faculty of Medicine
Tepper, B.J. (1998). 6-n-Propylthiouracil: a genetic marker for
University of Toronto
taste, with implications for food preference and dietary habits.
150 College Street
Am J Hum Genet 63, 1271–1276.
Tepper, B.J. (2008). Nutritional implications of genetic taste vari- Toronto, Ontario, M5S 3E2, Canada
ation: the role of PROP sensitivity and other taste phenotypes.
Annu Rev Nutr. E-mail: a.el.sohemy@utoronto.ca

También podría gustarte