ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img

Transient Receptor Potential Channels as Targets for Phytochemicals

View Author Information
Department of Pharmacology, Southern Illinois University School of Medicine, Springfield, Illinois 62702, United States
*E-mail: [email protected]. Phone: 217-545-2179. Fax: 217-545-0145.
Cite this: ACS Chem. Neurosci. 2014, 5, 11, 1117–1130
Publication Date (Web):June 13, 2014
https://doi.org/10.1021/cn500094a

Copyright © 2014 American Chemical Society. This publication is licensed under these Terms of Use.

  • Open Access

Article Views

7172

Altmetric

-

Citations

60
LEARN ABOUT THESE METRICS
PDF (1 MB)

Abstract

To date, 28 mammalian transient receptor potential (TRP) channels have been cloned and characterized. They are grouped into six subfamilies on the basis of their amino acid sequence homology: TRP Ankyrin (TRPA), TRP Canonical (TRPC), TRP Melastatin (TRPM), TRP Mucolipin (TRPML), TRP Polycystin (TRPP), and TRP Vanilloid (TRPV). Most of the TRP channels are nonselective cation channels expressed on the cell membrane and exhibit variable permeability ratios for Ca2+ versus Na+. They mediate sensory functions (such as vision, nociception, taste transduction, temperature sensation, and pheromone signaling) and homeostatic functions (such as divalent cation flux, hormone release, and osmoregulation). Significant progress has been made in our understanding of the specific roles of these TRP channels and their activation mechanisms. In this Review, the emphasis will be on the activation of TRP channels by phytochemicals that are claimed to exert health benefits. Recent findings complement the anecdotal evidence that some of these phytochemicals have specific receptors and the activation of which is responsible for the physiological effects. Now, the targets for these phytochemicals are being unveiled; a specific hypothesis can be proposed and tested experimentally to infer a scientific validity of the claims of the health benefits. The broader and pressing issues that have to be addressed are related to the quantities of the active ingredients in a given preparation, their bioavailability, metabolism, adverse effects, excretion, and systemic versus local effects.

SPECIAL ISSUE

This article is part of the TRPs as Probes and Medications for CNS Disorders special issue.

David Julius and colleagues from the University of California, San Francisco, cloned the receptor for the active ingredient in hot chili pepper, capsaicin, and named it as vanilloid receptor 1 (VR1) because capsaicin has a vanillyl moiety in its structure. (1) It was recognized that the VR1 had a sequence homology to a receptor cloned from a mutant fly (Drosophila melanogaster), in which the electroretinogram exhibited a transient response to continuous light; (2) therefore, it was renamed as transient receptor potential (TRP) Vanilloid 1 (TRPV1). (3, 4) Several TRP channels have been cloned, and some of them are considered as targets for active ingredients in botanicals. For example, TRP Ankyrin 1 (TRPA1), a receptor that carries sensory information from the periphery, is coexpressed with TRPV1, activated by the active ingredient in mustard, allyl isothiocyanate (AITC), (5, 6) and TRP Melastatin 8 (TRPM8), involved in sensing cold temperatures, is activated by menthol extracted from mint leaves. (7, 8) Other TRP channels activated by plant ingredients include TRPC6 by hyperforin, TRPV3 by incensole, and TRPM5 indirectly by glucose (sweet taste receptor-mediated increase in intracellular Ca2+). All the subunits of TRP channels that have been cloned so far have six transmembrane domains and a loop between domains five and six, which forms the pore. The channels have a stoichiometry of homo- or heterotetramers. Recently, high resolution structural studies using electron cryomicroscopy have confirmed the tetrameric structure. (9, 10)
Intense research is ongoing to identify the active ingredients in botanicals and their targets to explain the physiological effects they claim to exert. The active ingredient in turmeric is curcumin, which is claimed to be effective in conditions ranging from relieving flatulence to treating Alzheimer’s disease and cancer. However, the bioavailability of curcumin is very low to produce effects systemically either because it is not absorbed or because it is metabolized rapidly by the liver (first-pass metabolism). However, it is important to emphasize that these ingredients can cause significant effects locally in the gastrointestinal (GI) tract. Activation of TRP channels can modulate or promote the release of peptide hormones and neurotransmitters from the sensory and enteric nerve endings and from enteroendocrine cells. It is fascinating to learn that, following a specific type of Bariatric surgery (Roux-en-Y) that involves transposition of the ileum, where the ileum is directly connected to the stomach, unexpectedly, the glucagon-like peptide-1 (GLP-1) levels increased significantly. (11) It is inferred from this effect that when food is directly exposed to the ileum, certain ingredients are able to stimulate specialized cells in the lower GI tract (enteroendocrine cells) to cause release of GLP-1. In an intact GI tract, these ingredients may be degraded because of the high acidic environment (pH 2–3) of the stomach and by the gastric enzymes that are released during digestion. This further emphasizes the local effects of phytochemicals in the GI tract, although some of them are not well absorbed. The purpose of this Review is to provide scientific bases for the effects of plant products by identifying the phytochemicals and their TRP channel targets.

Phytochemicals That Activate Transient Receptor Potential Ankyrin (TRPA)

ARTICLE SECTIONS
Jump To

TRPA1 is the only identified member of this family, which is a Ca2+ permeable nonselective cation channel, predominately expressed in a population of sensory neurons that also express TRPV1. (5, 12) TRPA1 is activated by the phytochemicals such as allyl isothiocyanate (AITC), allicin, diallyldisulfide (DADS), cinnamaldehyde, methylsalicylate, Δ-9-tetrahydrocannabinoid (THC), and synthetic compounds such as icilin, acrolein, N-methylmaleimide (NMM), and (R)-(+)-[2,3-dihydro-5-methyl-3-(4-morpholinylmethyl)pyrrolo[1,2,3-de)-1,4-benzoxazin-6-yl]-1-apthalenylmethanone (WIN55,212-2). (6, 13-16) TRPA1 can be activated by multiple products of oxidative stress, which include hydrogen peroxide (H2O2), hydroxyalkenyl aldehyde (4-hydroxynonenal, 4-HNE), and 15-deoxy-Δ12,14-prostaglandin J2 (15d-PGJ2) (17, 18) (Figure 1). TRPA1 can also be activated by bradykinin (BK). (13) Recent studies show that TRPA1 is activated by methylglyoxal (MG). (19) MG is formed from triose phosphate during secondary glucose metabolism in hyperglycemic conditions. It is well-known that MG covalently modifies arginine, lysine, and cysteine residues and forms advanced glycation end products. (20) TRPA1 has been shown to be involved in various sensory processes, such as detection of noxious cold, mechanosensation, and inflammatory hyperalgesia. (5, 13, 16, 21-23)

Figure 1

Figure 1. TRPA1 agonists (phytochemicals, synthetic chemicals, and endogenous molecules). Allicin, 2-propene-1-sulfinothioic acid S-2-propenyl ester; allylisothiocyanate (AITC), 3-isothiocyanato-1-propene, is an organosulfur compound; cinnamaldehyde, (2E)-3-phenylprop-2-enal; Δ9-tetrahydrocannabinol, (−)-(6aR,10aR)-6,6,9-trimethyl-3-pentyl-6a,7,8,10a-tetrahydro-6H-benzo[c]chromen-1-ol; umbellulone, 1-isopropyl-4-methylbicyclo[3.1.0]hex-3-en-2-one; ligustilide, (3Z)-3-butylidene-4,5-dihydro-2-benzofuran-1(3H)-one; methyl salicylate, methyl 2-hydroxybenzoate; icilin, 1-(2-hydroxyphenyl)-4-(3-nitrophenyl)-3,6-dihydropyrimidin-2-one; N-methylmaleimide (oxidizing agent); 4-hydroxynonenal, 4-hydroxy-2-nonenal, an α,β-unsaturated hydroxyalkenal produced by lipid peroxidation, is an endogenous agonist; methylglyoxal, an aldehyde from pyruvic acid, acts both as an aldehyde and ketone, and reacts with free amino acids such as lysine, arginine and thiol groups of cysteine. MG is an endogenous agonist.

TRPA1 can be activated by three different mechanisms, including a mechanism of covalent modification of cysteine residues, which is unique among ion channel activation mechanisms: (1) AITC, allicin, DADS, acrolein, and NMM activate the channel by covalent modification of cysteine residues in the cytoplasmic N-terminals; (2) THC and WIN55,212-2 activate the channel possibly by binding to a site; and (3) BK by activating phospholipase C (PLC). Simultaneous mutations of C619, C639, and C663 significantly reduced NMM- and AITC-induced current. It was further demonstrated that additional mutation of K708 prevented the activation by AITC but THC could still activate the channel. (23) Cysteine residues are involved in covalent modification, yet the membrane current responses induced by TRPA1 agonists such as AITC and NMM are readily reversible. (24)
The active ingredients in cinnamon (Cinnamomum zeylanicum), which belongs to the family Lauraceae, are cinnamaldehye, cinnamyl alcohol, and cinnamyl acetate. Cinnamon is a sweet-smelling spice obtained from the bark of the tree. In early days it was used in perfumes, as an appetite stimulant, and to flavor wines. It is considered to improve digestion and acts as an aphrodisiac and is found to be effective in treating sore throat and common cold.
Cinnamaldehyde ((2E)-3-phenylprop-2-enal) is a TRPA1 agonist (Figure 1, EC50 = 100 μM). (13) The pungency of cinnamon, when it comes in contact with the tongue, is due to its ability to activate TRPA1 expressed at the nerve terminals. Further, the activation of TRPA1 can cause the release of vasoactive peptides, such as calcitonin gene-related peptide (CGRP) and substance P (SP) from the nerve terminals. It is intriguing that fibers that carry pain sensation also innervate the blood vessels, although the blood vessels are considered to be insensate. (25) It is likely that the vasoactive substances released from the nerve terminals have beneficial effects on the cardiovascular functions. Activation of these receptors in the nerve terminals innervating the GI tract sends signals to satiety centers and releases neuropeptides/neurotransmitters locally. It has been shown that cinnamon can decrease blood glucose levels in type 2 diabetes. (26, 27) Diabetic animals treated with cinnamon showed decrease in blood glucose levels, which could be brought about by the release of incretins (glucose-dependent insulinotropic hormone (GIP) and GLP-1) and insulin release caused by activation of TRPA1 receptors. (19)
Garlic (Allium sativum) belongs to the family Alliaceae. There are several claims that consumption of garlic imparts good health. Beneficial effects of garlic in fighting common cold, sore throat, and cough have been reported. (28, 29) Allicin, the active ingredient in garlic, is an organosulfur compound found to have potent antibacterial and antifungal properties (30) (Figure 1). Allicin activates TRPA1 and TRPV1. (15) It is important to note that allicin has a very short half-life (1–5 s), because it rapidly decomposes. When garlic is crushed, the pungent smell is due to formation of allicin from alliin by the enzyme allicinase. When allicin is degraded, it forms 2-propenesulfenic acid, which can bind to free radicals. Other sulfur compounds present in garlic are ajoene, allyl sulfides, and vinyldithiins.
Anticancer effects of ingredients in garlic have been shown in cell lines and in animal experiments. (31) Local application of ingredients in garlic can prevent certain forms of skin cancer. (32) The anticancer effects of ingredients in garlic could be due to a local effect of the phytochemicals in the gastrointestinal tract activating TRPA1 and promoting Ca2+ influx, rather than being absorbed and acting systemically. Excessive Ca2+ flux leads to cell death.
Mustard belongs to the family Brassicaceae, genus Brassica and species alba (yellow mustard); or genus Snapis and species nigra (black mustard). Mustard seed contains several ingredients, such as glucosinolates (sinigrin), that can be broken down by the enzyme myrosinase to yield isothiocyanates. The active ingredient in pure mustard oil, AITC, activates TRPA1 and is involved in several functions (see cinnamon and garlic) (Figure 1). While eating mustard-laced food, the effect is more of an olfactory sensation (smell) rather than gustatory sensation (taste). Derivatives of AITC have been used as a war gas. Nitrogen mustard (mechlorethamine) is used as an anticancer agent. As discussed earlier, TRPA1 is a highly Ca2+-permeable channel, the activation of which can cause neuropeptide/neurotransmitter release and can mediate intracellular Ca2+-induced cellular functions.
AITC also is the active phytochemical in horseradish (Armoracia rusticana), which belongs to the family Brassicaceae. It is considered to be pungent; the pungency is due to activation of TRPA1 expressed at the nerve terminals. Wasabi is a common condiment in Japanese cuisine, which has high levels of AITC. When taken orally, it is considered as a remedy for sinusitis, sore throat, and nasal congestion. It is considered to be anthelmintic and bactericidal. Thiopropanal S-oxide is the pungent ingredient in onion (Allium cepa) which belongs to the family Alliaceae. Similar to allicin and AITC, isothiocyanate and thiopropanal S-oxide activate the ion channels TRPA1. (33) AITC activates the channel with an EC50 of 33 μM. (13) When ingested orally, the bioavailability of AITC is high and is considered as a cancer chemopreventive compound. (33)
Curcumin ((E,E)-1,7-bis(4-hydroxy-3-methoxyphenyl)-1,6-heptadiene-3,5-dione), obtained from turmeric (Curcuma longa), which belongs to the family Zingiberaceae, is an activator of TRPA1. (34) The spice comes from the root of the plant and has a bright-yellow color. There are several claimed effects of curcumin, which include anti-inflammatory, antioxidant, anticancer, antidiabetic, antimicrobial, and so forth. (35) It has been claimed to be effective in a wide variety of conditions, including flatulence, jaundice, menstrual pain, toothache, and colic. The bioavailability of curcumin is very low. In a study, an oral dose of 8 g resulted in blood levels of around 200 ng. (35) However, efforts are being made to improve curcumin bioavailability by various stable preparations, including packaging curcumin in lipid nanoparticles. Longvida is a solid lipid curcumin particle (SLCP) that has a high bioavailability. As discussed earlier, it is possible that turmeric could produce its effects locally by direct contact with the cells in the lumen of the GI tract.
Cystic fibrosis is a condition caused by mutations in the cystic fibrosis transmembrance conductance regulator (CFTR). Curcumin has been shown to be effective in the disease caused by ΔF508 mutation, which results in the production of misfolded CFTR protein. (36-38) Curcumin activates TRPA1, which is expressed in the bronchial mucosa and relieves symptoms of cystic fibrosis.
The phytochemical umbellulone is obtained from Umbellularia californica, which belongs to the family Lauraceae (Figure 1). The tree is called “headache tree” because the vapors from the tree can cause severe headache. Umbellulone is a reactive molecule that binds to cysteine residues in TRPA1, thereby activating the receptor (EC50 = 11.6 μM). The headache may be due to activation of TRPA1 in the trigeminal system. Stimulation of sensory nerve terminals causes neuropeptide release. Since CGRP has been shown to play an important role in migraine type headaches by causing vasodilation of the meningeal vessels, it is likely that umbellulone causes headache by causing CGRP release by activating TRPA1. (39)
Ligustilide, a dihydrophthalide, is a reactive molecule obtained from Angelica acutiloba, which belongs to the family Apiaceae (Figure 1). Ligustilide can bind to thiol groups, and this property could be responsible for activating TRPA1 (EC50 = 44 μM). As this plant ages, it produces dehydroligustilide (EC50 = 539 μM), but this is an antagonist of TRPA1 at lower concentrations (IC50 = 23 μM). The specificity and the reactive nature of these molecules were demonstrated by mutating specific residues required for TRPA1 activation by electrophilic and reactive TRPA1 agonists. Celery contains ligustilide, which could activate TRPA1 and bring about its beneficial gustatory effects. (40)
Paclitaxel obtained from the Pacific yew (Taxus brevifolia), which belongs to the family Taxaceae, is being used to treat certain forms of cancer (Figure 2). One of the side effects of paclitaxel is peripheral neuropathy that can be explained by its ability to activate TRPA1, which mediates tactile and cold allodynia. (41)

Figure 2

Figure 2. Other TRP channel activators. Hyperforin, (1R,5S,6R,7S)-4-hydroxy-5-isobutyryl-6-methyl-1,3,7-tris(3-methyl-2-buten-1-yl)-6-(4-methyl-3-penten-1-yl)bicyclo[3.3.1]non-3-ene-2,9-dione (TRPC6); bisandrographolide, 3-{(E)-2-[6-hydroxy-5-(hydroxymethyl)-5,8a-dimethyl-2-methylenedecahydro-1-naphthalenyl]vinyl}-5-{6-hydroxy-5-(hydroxymethyl)-5,8a-dimethyl-2-methylene-1-[2-(2-oxo-2,5-dihydro-3-furanyl)ethyl]decahydro-1-naphthalenyl}-2(5H)-furanone (TRPV4); paclitaxel, (2α,5β,7β,10β,13α)-4,10-diacetoxy-13-{[(2R,3S)-3-(benzoylamino)-2-hydroxy-3-phenylpropanoyl]oxy}-1,7-dihydroxy-9-oxo-5,20-epoxytax-11-en-2-yl benzoate (TRPA1, TRPV4); dicentrine, (7aS)-10,11-dimethoxy-7-methyl-6,7,7a,8-tetrahydro-5H-[1,3]benzodioxolo[6,5,4-de]benzo[g]quinoline (TRPA1).

Δ(9)-Tetrahydrocannabinol (THC) is a psychoactive compound in Cannabis sativa, which belongs to the family Cannabaceae (Figure 1). As described earlier TRPA1 is activated by covalent modification of cysteine residues. However, when the cysteine residues were mutated, the activation by NMM was abolished, but THC and WIN55,212-2 could still activate the channel possibly by binding to a site. (22) The phytochemicals in cannabis, cannabichromene and cannabigerol, activate TRPA1 with an EC50 of 60 nM and 3.4 μM, respectively. Cannabidiol acid was least potent (EC50 ∼ 12 μM) (42) (Table 1).
Table 1. Phytochemicals That Modulate TRP Channels
plant name phytochemical TRP target EC50 (μM)a ref
Cinnamomum zeylanicum cinnamaldehyde TRPA1 6.8 6
      61 13
Allium sativum allicin TRPA1 7.5 22
  DADS TRPA1 192 22
Brassica alba allyl isothiocyanate TRPA1 64.5 23
Snapis nigra     22 13
      11 6
Curcuma longa curcumin TRPA1 ND 34
Umbellularia californica umbelluline TRPA1 11.6 39
Angelica acutiloba ligustilide TRPA1 44 40
dehydroligustilide TRPA1 539 40
Taxus brevifolia paclitaxel TRPA1 ND 41
    TRPV4 ND 41
Cannabis sativa tetrahydrocannabinol TRPA1 12 6
    TRPM8 0.1 (ant) 42
  cannabidiol TRPA1 12 42, 119
    TRPV1 3.2 42
    TRPM8 0.1 (ant) 42
    TRPV3 3.7 42
  cannabichromene TRPA1 0.06 42
  cannabigerol TRPM8 0.1 (ant) 42
    TRPA1 3.4 144
  cannabidivarin TRPV4 0.9 144
  tetrahydrocannabivarin TRPV3 3.7 144
    TRPV4 6.4 144
Lindera megaphylla dicentrine TRPA1 ND 43
Nicotiana tabacum nicotine TRPA1 ∼10 44
Zingiberaceae aframomum melegueta linalool TRPA1 117 45
    TRPM8 6700 116
  α-sanshool TRPA1 69 45
    TRPV1 1.1  
  shogaol TRPA1 11.2 45
    TRPV1 0.2  
  paradol TRPA1 71 45
    TRPV1 1.8  
Hypericum perforatum hyperforin TRPC6 0.7 (Na+) 48
      1.2 (Ca2+)
Mentha longifolia menthol TRPM8 80 8
    TRPA1 68 (ant) 142
Eucalyptus globulus eucalyptol TRPM8 7700 66
Tripterygium wilfordii triptolide TRPP ND 67
Capsicum annuum capsaicin TRPV1 0.71 1
      0.3 119
Euphorbia resinifera Euphorbia resinifera TRPV1 0.04 1
Ocimum basilicum eugenol TRPV1 ND 13, 115
Cinnamonium tamala   TRPV3 ND 123
Artemisia dracunculus   TRPA1 262 13, 116
    TRPM8 ND 13, 116
Piper nigrum piperine TRPV1 38 117
Cinnamomum camphora camphor TRPV1 4500 120, 142
Rosmarinus officinalis   TRPA1 660 (ant) 120
      68 (ant) 142
Euodia ruticarpa evodiamine TRPV1 0.86 121
Zingiber officinale gingerols TRPV1 ND 13, 122
    TRPA1 ND 13
Origanum vulgare thymol TRPV1 ND 123
  carvacrol TRPV3 ND 123
    TRPA1 ND 123
Thymus vulgarism thymol TRPV3 ND 123
Tasmannia lanceolata polygodial TRPV1 ND 124
Vanilla planifolia vanillin TRPV1 ND 123, 126
    TRPA1 ND 123
Vernonia tweedieana α-spinasterol TRPV1 40 (ant) 127
Boswellia thurifera incensole TRPV3 16 143
Andrographis paniculata bisandrographolide TRPV4 0.87 159
a

(Na+), sodium flux; (Ca2+), calcium flux; (ant), antagonist.

Dicentrine is a naturally occurring aporphine type isoquinoline alkaloid, isolated from the root Lindera megaphylla Hemsl., which belongs to the Lauraceae family (Figure 2). In animal models, dicentrine induced antinociceptive effects. Cinnamaldehyde-induced nocifensive behavior was abolished by dicentrine, but not the capsaicin-induced nocifensive behavior. Based on these studies, it is proposed that the dicentrine effect may involve interaction with TRPA1 channels. (43)
Nicotine is obtained from Nicotiana tabacum of the Solanaceae family. It brings about its stimulatory actions by activating neuronal nicotinic acetylcholine receptors. However, it also produces irritation while smoking, chewing, or snorting. It has been shown that irritation is caused by the activation of TRPA1. Nicotine activates TRPA1 in lower concentrations (EC50 ∼ 10 μM), but inhibits at higher concentrations (>1 mM). (44)
Extracts of Sichuan and melegueta peppers evoke pungent sensations that are mediated by different alkylamides, such as sanshool and shogaol; both activate TRPA1 and TRPV1 channels. Linalool, a terpene in Sichuan peppers, is able to activate TRPA1 but not TRPV1 (45) (Table 1).

Phytochemicals That Activate Transient Receptor Potential Canonical (TRPC)

ARTICLE SECTIONS
Jump To

TRPC channels have been classified as TRPC1, TRPC2, TRPC3/6/7 and TRPC4/5 on the basis of structural similarities and functions. (46) The channel is formed as a homo- or heterotetramer. The activation mechanism of TRPC channels is not fully clarified. These channels are associated with G-protein coupled receptors and activation of G-protein coupled receptor results in transactivation of TRPC channels and facilitates their openings. (47)
St. John’s wort (Hypericum perforatum) extract is used as an antidepressant, and the mechanism of action is still elusive. One of the active ingredients in the extract has been found to be hyperforin, a bicyclic polyprenylated acylphloroglucinol compound (Figure 2). Hyperforin has been shown to activate TRPC6. (48) In general, antidepressants are selective serotonin and norepinephrine uptake inhibitors, thereby increasing the levels of serotonin and norepinephrine. The neurotransmitter levels can also be increased by promoting their release by causing Ca2+ influx at the nerve terminals. Generally, neurotransmitter release occurs in response to an action potential arriving at the nerve terminal and activating voltage-gated Ca2+ channels. It is becoming increasingly apparent that Ca2+ permeable TRP channels expressed at the presynaptic terminals can cause transmitter release and modulate synaptic transmission, independent of action potentials. TRPV1 and TRPA1 expressed in the presynaptic terminals of the sensory neurons can cause neurotransmitter release. (49-52) The effect of hyperforin could be due to its activation of TRPC6 expressed in central neurons. The effect appears to be specific because TRPC3 is unaffected by hyperforin. TRPC6, like most TRP channels, is a nonspecific cation channel that has a high Ca2+ permeability. Hyperforin also exhibits neurotrophic effects leading to axonal sprouting and neurite extension. In confirmation of this effect, neurons from TRPC6 overexpressing animals exhibit enhanced dendritic growth and synapse formation that may play a role in learning and memory formation. (53)

Phytochemicals That Activate Transient Receptor Potential Melastatin (TRPM)

ARTICLE SECTIONS
Jump To

There are eight members in TRPM family. There are no ankyrin domains in the N-terminus of these channels. These channels are Ca2+ and Mg2+ permeable, and the permeability to Ca2+ ranges from impermeable (TRPM4 and TRPM5) to significantly Ca2+ permeable (TRPM6 and TRPM7). (54)

TRPM5

TRPM5 is a channel activated by increases in intracellular Ca2+. TRPM5 activation by sweet tastants is by an indirect mechanism of causing an increase in intracellular Ca2+ levels by the phytochemicals activating the sweet-taste receptor. Several structurally diverse phytochemicals have been shown to activate the sweet-taste receptor. But the degree of sweetness differs; glucose is less sweet in comparison to sucrose, which is a disaccharide formed by the combination of fructose and glucose. Fructose is the sweetest (73% sweeter than glucose). The plant product stevioside, obtained from Stevia rebaudiana, activates the sweet receptors. Interestingly, although the whole-plant extract can be used as a sweetening agent, it has been shown to cause infertility. However, the pure ingredient from this plant, rebaudioside A, is devoid of this action. (55)
The type 2 taste receptor (T2R) is alpha-gustducin, which is a sweet- and bitter-taste receptor. A knockout of the alpha-gustducin gene causes animals to lose both sweet and bitter taste sensations. T2Rs sense bitter taste. For umami and sweet taste perception, taste receptors type 1 (T1R1), type 2 (T1R2), and type 3 (T1R3) form heterodimers. T1R1 and T1R3 form a complex to sense the umami taste. T1R2 and T1R3 form a complex to taste sweetness. The signal transduction involves the activation of the G-protein-coupled receptor alpha-gustducin, which is coupled to phospholipase Cβ 2 (PLCβ 2). Activation of PLCβ 2 promotes the hydrolysis of PIP2 to form IP3 and DAG. IP3 releases Ca2+ from intracellular stores and activates TRPM5, which depolarizes the cell and causes ATP release. ATP acts as a neurotransmitter and mediates signal transduction. (56)
It is becoming evident that the receptors that sense sweet-taste on the tongue are also expressed throughout the GI tract and act as chemosensors. Taste receptors are present in the cells lining the stomach, pancreas, and enteroendocrine cells of the GI tract. Stimulation of brush border cells with tastants releases GLP-1 and peptide YY (PYY). Both alpha-gustducin and TRPM5 receptors are predominately expressed in these cells. It is possible that phytochemicals are able to activate these receptors and promote release of neuropeptides, neurotransmitters and hormones. (57-60)

TRPM8

Mint (Mentha longifolia) belongs to the family Lamiaceae, and peppermint belongs to the species piperita. The decoction of mint leaves is used for stomach aches and for some painful conditions. The active ingredient is menthol (Figure 3); it is used as an ingredient in various balms that are used to relieve pain.

Figure 3

Figure 3. TRPM8 agonists. Menthol, (1R,2S,5R)-2-isopropyl-5-methylcyclohexanol; eucalyptol, 1,3,3-trimethyl-2-oxabicyclo[2.2.2]octane; icilin, 1-(2-hydroxyphenyl)-4-(3-nitrophenyl)-3,6-dihydropyrimidin-2-one.

Transient receptor potential melastatin 8 (TRPM8), previously known as menthol and cold receptor 1 (CMR1), is a Ca2+ permeant nonspecific cation channel, which is expressed in a subpopulation of primary afferent neurons. TRPM8 is activated by cold (<25 °C), phytochemicals such as menthol and eucalyptol, and the synthetic chemical icilin. Activation of TRPM8 induces a cool/soothing sensation. (7, 8) TRPM8 expressed at the central terminals modulates synaptic transmission. (61, 62) TRPM8 (earlier identified as Trp-p8) is upregulated in prostate cancer and is involved in urinary bladder functions, which broadens the horizon of the involvement of TRPM8 in other physiological and pathophysiological conditions. (63)
Generally, phosphorylation enhances the activity of ion channels. However, it has been demonstrated that a functional downregulation of TRPM8 occurs when PKC is stimulated resulting in an inhibition of TRPM8-mediated channel activity, in contrast to TRPV1, which is robustly potentiated by PKC activation. These effects are due to dephosphorylation of TRPM8 by activation of protein phosphatases. (61)
The activation of TRPM8 sends the cool and soothing sensation to alleviate pain. Therefore, it is expected to be upregulated by phosphorylation in inflammatory conditions. Paradoxically, phosphorylation downregulates TRPM8, thereby compromising the much needed cool and soothing sensation.
It has also been shown that mentholated cigarette smoke exerts a cool and soothing sensation while inhaling; because of this, mentholated cigarettes may encourage the smoking habit. Recently, the United States Food and Drug Administration has issued a warning that mentholated cigarettes are more addictive. (64) Further, menthol can directly interact with the nicotinic acetylcholine receptor and inhibit its function, a mechanism that may explain the reason for smoking a greater number of cigarettes to get the same effect, thereby increasing the addictive potential of nicotine. (65)
Eucalyptol is obtained from the leaves of Eucalyptus globulus in an oil form (Figure 3). It has the structure of a cyclic ether and a monoterpenoid. It has a smell resembling that of camphor. It is added as one of the additives in cigarettes. Eucalyptol is a TRPM8 agonist (EC50 = 7.7 mM). (66)

Phytochemicals That Activate Transient Receptor Potential Polycystin (TRPP)

ARTICLE SECTIONS
Jump To

The TRPP family is made up of three channel members, namely, TRPP1, TRPP2, and TRPP3. TRPP1 is an ion channel, which is considered to be involved in polycystic kidney disease. The disease is characterized by the formation of multiple cysts, hence the name polycystic kidney disease, eventually leading to kidney failure. In this disease, cysts are also found in liver, pancreas, and other inner surfaces covered by tubular epithelial cells. In tubular epithelial cells, ciliary action transduces a mechanical stimulus and opens a Ca2+ permeable ion channel, such as polycystin-2 (PC2 or TRPP1), and increases the intracellular Ca2+ levels and causes cell cycle arrest. Mutations in TRPP and/or the associated protein, polycystin 1 (PC1), result in autosomal dominant polycystic kidney disease (ADPKD).
Triptolide, a diterpene (Figure 2) obtained from Tripterygium wilfordii, induces Ca2+ influx in tubular epithelial cells and controls their proliferation, resulting in reduced cyst formation and alleviation of symptoms associated with kidney damage in a murine model of ADPKD. (67)

Phytochemicals That Activate Transient Receptor Potential Vanilloid (TRPV)

ARTICLE SECTIONS
Jump To

There are six members in the TRPV family, the name is derived by the activation of TRP Vanilloid 1 (TRPV1) by molecules consisting of a vanillyl moiety. The other members of this family are TRPV2, TRPV3, TRPV4, TRPV5, and TRPV6. (68-70) While comparing the Ca2+ permeability, it has become clear that TRPV5 and TRPV6 channels are purely Ca2+ permeable. They form homo- or heterotetrameric structures. (71) The high resolution structure of TRPV1 has been recently published, aided by electron cryomicroscopy. (9, 10)

TRVP1

TRPV1, formerly known as vanilloid receptor 1 (VR1), is a nonselective cation channel with high Ca2+ permeability, which is expressed predominantly in a population of small-diameter sensory neurons. It functions as a polymodal receptor in the peripheral sensory nerve terminals and modulates synaptic transmission at the first sensory synapse. (1, 49, 51, 72-74) A recent study has shown its expression in the dorsal horn inhibitory interneurons. (75) Capsaicin has been shown to modulate synaptic transmission in other brain regions. (76-80) Activation of TRPV1 results in two functional components: (1) sending the impulses to the brain by generating an action potential at the nerve endings (afferent function) and (2) releasing vaso/neuroactive substances by virtue of its Ca2+ permeability, such as histamine, bradykinin, CGRP, and SP, from the peripheral nerve terminals (efferent action).
TRPV1 is activated by heat (>42 °C) and phytochemicals such as capsaicin, resiniferatoxin (RTX), tinyatoxin (TNX), camphor, carvacrol, and thymol. It is activated by endogenous ligands such as protons, anandamide, arachidonic acid metabolites, and N-arachidonyl dopamine (NADA). (1, 27, 49, 54, 68, 81-86)
Although TRPV1 is considered mainly to be involved in thermal sensory perception, its distribution in regions that are not exposed to its activation temperature ranges raises the possibility of its involvement in other functions. TRPV1 can be detected using RT-PCR and radioligand binding throughout the neuroaxis, and the identification of specific ligands such as NADA in certain brain regions further suggests its roles in the CNS. (51, 75, 87-90) TRPV1 is present in the smooth muscles of the blood vessels and bronchi, where activation of the receptor leads to vasodilation by releasing CGRP, acetylcholine, or nitric oxide from nerve terminals and bronchoconstriction by promoting Ca2+ influx, respectively. (91-93) TRPV1 is found in the nerve terminals, supplying the urinary bladder and the urothelium, indicating a role in urinary bladder function, such as micturition. (94, 95) Interestingly, TRPV1 is also involved in the regulation of body temperature. Subcutaneous injection of capsaicin decreases the body temperature by 2–3 °C. TRPV1 antagonists increase the body temperature to the same extent. (96, 97)
TRPV1 has emerged as a potential target for developing analgesics. Potent TRPV1 antagonists have been developed and shown to be effective in alleviating pain in several animal models. Unfortunately, development of hyperthermia following their administration has halted the clinical trials. (94) However, findings suggest that certain compounds may be devoid of the effect of elevating core body temperature. (98)
Capsaicin (8-methyl-N-vanillyl-6-nonenamide) (Figure 4), an active ingredient in hot chili pepper (Capsicum annuum or frutescens), which belongs to the family Solanaceae. The nonpungent bell peppers belong to the species annuum, and hot peppers belong to the species frutescens.

Figure 4

Figure 4. TRPV1 agonists. Vanillin, has the vanillyl moiety, that is essential for activating TRPV1 channels; Capsaicin, (E)-N-[(4-hydroxy-3-methoxyphenyl)methyl]-8-methyl-6-nonenamide, has a vanilloid and an aceyl moiety; dihydrocapsaicin, the structure of which is a 6,7-dihydro derivative of capsaicin; resiniferatoxin, has a complex structure, but shares a homovanillyl group, which is necessary for the activity of all vanilloids; eugenol, 2-methoxy-4-(2-propenyl)phenol and is a member of the allylbenzene class of chemical compounds; Cannabidiol, 2-[(1R,6R)-6-isopropenyl-3-methyl-3-cyclohexen-1-yl]-5-pentyl-1,3-benzenediol; anandamide or arachidonylethanolamide or arachidonic acid N-(hydroxyethyl)amide consists of the acyl moiety and is an edogenous ligand of TRPV1 and cannabinoid receptor 1 (CB1).

The hotness of chili peppers is due to the chemical content of capsaicin. Purified capsaciin activated TRPV1 with a EC50 of 711.9 nM. (1) Other capsinoids such as capsiate, dihydrocapsiate, and nordihydrocapsiate have been isolated and purified. The degree of hotness can be quantified by using the Scoville scale. A Scoville unit is the “number” of times the alcoholic extract has to be diluted to lose the pungency. A sweet bell pepper has a score of 0 Scoville units and the Habanero, Savina, and Naga Jolokia (ghost peppers) are calibrated to have scores of 400 000, 600 000, and 1 000 000 Scoville units, respectively. For example, ghost pepper extract has to be diluted one million times to lose its pungency. Capsaicin and dihydrocapsaicin have the Scoville scores of 16 000 000 and 15 000 000, respectively. Shogaol from ginger has the score of 160 000 units, piperine from pepper has the score of 100 000 units, and gingerol from ginger has a score of 60 000 units. Commercially available pepper spray has the score of 2 500 000 units, and police-grade pepper spray has the score of 5 500 000 Scoville units.
When capsaicin binds to its receptor, the ion channel opens, but when constantly activated, the receptor enters a desensitization state. On the other hand, sensitization is a phenomenon, where the receptor activity is enhanced by phosphorylation. Overexpression and overactivation of TRPV1 is observed in various painful conditions. Topical capsaicin application has been useful to treat conditions such as arthritis, diabetic peripheral neuropathy, shingles, and psoriasis by exerting a local effect. (85, 86, 99, 100) The mechanism of pain relief has been proposed to be due to desensitization of the receptor or degeneration/ablation of the nerve terminals. (101)
Altered expression of TRPV1 is found in cancers involving prostate, bladder, pancreas, tongue, skin, liver, and colon. Capsaicin can induce its effects by causing Ca2+ influx through TRPV1 overexpressed in cancerous cells, which can lead to cell death by apoptosis or necrosis. (85)
In diabetes, consuming a hot chili pepper containing meal showed a decrease in the amount of insulin required to combat the postprandial increase in glucose. TRPV1 may play a role in this effect. (102, 103) Capsaicin has been reported to increase oxygen consumption and thermogenesis, which might lead to weight loss. There are studies to support a potential neurogenic mechanism by which TRPV1-sensitive sensory neurons may regulate energy and fat metabolism. Capsaicin prevents adipogenesis by apoptotic mechanism. Capsaicin or N-oleoylethanolamide, an endogenous ligand of TRPV1, reduced food intake by conveying information through the vagus nerve and affecting satiety centers. TRPV1 knockout animals, when fed a high-fat diet, although the food intake was the same, they gained less weight as compared to their wild-type counterparts. (104, 105)
Although there is no direct evidence that hot chili pepper containing spicy food increases acid secretion in the stomach, in certain conditions, such as gastroesophageal reflux disease (GERD), increased expression of TRPV1 in the esophagus can induce a burning sensation. (106) Capsaicin has been shown to worsen the condition in patients with irritable bowel syndrome (IBS) and Crohn’s disease. Activation of TRPV1 has been shown to cause the release of gastric acid in the stomach, but other studies have reported otherwise. (107)
Urinary bladder hyperreflexia is a condition that has been shown to be related to overexpression of TRPV1 in the nerve terminals innervating the bladder. Excretion of capsaicin through the kidneys can accumulate in the bladder and exert an effect on the urinary bladder. (85)
Resiniferatoxin (RTX) and tinyatoxin (TNX) are the most potent among all the known natural, synthetic, and endogenous agonists of TRPV1. These pure chemicals obtained from a cactus-like spurge Euphorbia resinifera/poissonii. Purified RTX activates TRPV1 with an EC50 of 39.1 nM. (1) In fact, RTX can maximally activate single channel TRPV1 currents in picomolar ranges. (108) RTX/TNX have the Scoville scores of 16 000 000 000 and 5 300 000 000, respectively. RTX, a phorbol related diterpene (resiniferonol 9,13,14-orthophenylacetate 20-homovanillate), has a complex structure with a phorbol and a vanillyl moiety (Figure 4). Its ultrapotency was thought to be due to the phorbol moiety, which could activate protein kinase C (PKC) and promote phosphorylation of TRPV1. This notion was abandoned because of the higher concentrations of RTX required to activate PKC. The tritiated form ([3H]RTX) has been used as a tool in ligand-binding assays. (68, 109) Binding of capsaicin and RTX to TRPV1 involves amino acid residues, which have been shown to reside within N- and C-cytosolic and transmembrane domains of the channel. (110-112) RTX could induce nerve terminal ablation by sustained Ca2+ influx and prevent nociceptive transmission. Intrathecal administation of RTX provides a long-lasting pain relief. (51, 113, 114) A clinical trial is ongoing to determine the effectiveness of intrathecal administation of RTX in debilitating terminal cancer pain conditions (NCT00804154). Intravesicular irrigation of RTX containing solution has yielded a significant improvement in urinary bladder hyperreflexia by its ability to ablate TRPV1 expressing nerve terminals in the urinary bladder. (95) Resiniferatoxin and tinyatoxin are used as pesticides.
Basil (Ocimum basilicum) belongs to the family Lamiaceae. Basil and oregano have the ingredient β-caryophyllene, a natural bicyclic sesquiterpene, which is an agonist of cannabinoid receptor 2 (CB2) and has been shown to be involved in anti-inflammatory actions. Eugenol, a phenylpropene and an allyl chain-substituted guaiacol is one of the active ingredients in basil and clove (Figure 4). Other ingredients include citral that provides the citrus smell to basil, camphene in the African blue basil, and anethole in licorice and basil. Eugenol has anti-inflammatory properties by blocking the cyclooxygenase enzyme. Eugenol activates TRPV1 and TRPV3 (115) (Table 1).
Bay leaves (Cinnamonium tamala) belong to the family Lauraceae. The active ingredients include β-caryophyllene, eugenol, and linalool, a naturally occurring terpene alcohol found in many flowers and spice plants. As discussed above, eugenol can activate TRPV1 and TRPV3 ion channels and linalool can activate TRPA1. (116)
Black pepper (Piper nigrum) belongs to the family Piperaceae. The active ingredients have been isolated. The main ingredient is piperine. Other alkaloids present in black pepper include chavicine and piperidine (Figure 3). Piperine activates TRPV1. (117) The pungency of pepper is due to the alkaloid piperine and is quantified to have a score of 100 000 Scoville units as compared to the ghost pepper, which has a score of 1 000 000 Scoville units. It is also used to treat conditions such as sore throat and bronchitis. It improves digestion and acts as a carminative. Piperine is known to inhibit the liver metabolizing enzyme CYP3A4, thereby increasing the bioavailability of other drugs. Piperine has been shown to significantly increase the bioavailability of curcumin by interfering with its metabolism. (118)
As discussed earlier, several ingredients in Cannabis sativa can activate TRP channels. (119) Cannabidiol (CBD) does not exhibit any psychotropic effects. Recently, CBD has gained attention because of its effectiveness in treating refractory epilepsies in children. CBD and other active ingredients in cannabis are considered as activators of TRPV1 (Figure 4). CBD activates TRPV1 with EC50 of 3.2 μM as compared to activation by capsaicin (EC50 of 0.3–0.7 μM). (1, 119)
Camphor, a terpenoid, is a transparent solid obtained from an evergreen tree Cinnamomum camphora. Another source of camphor is from dried rosemary (Rosmarinus officinalis). Camphor is an activator of TRPV1. (120)
Clove (Eugenia caryophyllis or Syzgium aromaticum) belongs to the family Myrtaceae. Clove oil is commonly used to treat toothaches and used as a local anesthetic in dental procedures. The principal active ingredients in clove are eugenol and salicylic acid. The effects of both of these active ingredients are useful in painful conditions. Eugenol activates TRPV1 and TRPV3 channels. (115)
Evodiamine, an active ingredient from the Rutadeae family of plants (Euodia ruticarpa), is a TRPV1 agonist. It is being used as a dietary supplement; it has been shown to induce thermogenesis. (121)
Ginger (Zingiber officinale) belongs to the family Zingiberaceae. The active ingredients in ginger are gingerols, shogaols, and zingiberene. They exist in different forms (6, 8, 10) depending on the alkyl carbon chain. The pungency is quantified to be 60 000 Scoville units. Structurally, it is related to capsaicin and piperine. Heat converts gingerol to zingerone. When gingerol is dehydrated, it is converted to a more potent compound, shogaol (EC50 = 0.2 μM; 160 000 Scoville units). All these ingredients can activate TRPV1 and TRPA1. (122) The alcoholic extract has been shown to possess antioxidant properties.
Oregano (Origanum vulgare) belongs to the family Lamiaceae. It contains several phytochemicals, such as thymol, carvacrol, and rosmarinic acid. The antioxidant properties are stronger than those of synthetic antioxidants, such as butylated hydroxytoluene (BHT) and butylated hydroxyanisole (BHA). Thymol and carvacrol can activate TRPV1 and TRPV3 ion channels (123) (Table 1).
Polygodial, a drimane-type sesquiterpene dialdehyde, is an active ingredient obtained from the Dorrigo pepper (Tasmannia lanceolata). Polygodial activates TRPV1 and TRPA1. (124, 125)
Tarragon (Artemisia dracunculus) is a perennial herb that belongs to the family Asteraceae. The active ingredient that is responsible for the characteristic taste is considered to be cis-pellitorin. The ingredients in tarragon oil include methyl eugenol (36%) and methyl chavicol (16%). Eugenol is a TRPV1 and TRPV3 channel agonist. (115)
Thyme (Thymus vulgarism) belongs to the family Lamiaceae. The major ingredient in thyme oil is thymol; other ingredients are borneol, linalool, myrcene, and p-cymene. It is incorporated as an antiseptic in mouthwash and in toothpaste. Thymol and linalool have been shown to activate TRPV3 and TRPA1, respectively. (45, 116, 123)
Vanilla (Vanilla planifolia) belongs to the family Orchidaceae (Figure 4). The smell is due to the active ingredient vanillin, which is structurally related to eugenol or guaiacol. It is used for flavoring ice creams, confectionaries, tobacco, beverages, and so forth. The natural flavor is due to vanillic aldehyde, but the artificial flavor ethylvanillin is as potent as the natural vanilla. It has mild CNS effects, is regarded as an aphrodisiac, and is useful to treat impotence. Vanillin can activate TRPV1 and TRPV3 ion channels. (126)
Most phytochemicals have been shown to activate TRP channels. The active phytochemical in the dichloromethane fraction from the leaves of the medicinal plant Vernonia tweedieana that belongs to the family Asteraceae was identified to be α-spinasterol, which acts as a potent antagonist of TRPV1. The antagonistic effects were demonstrated by the displacement of tritiated RTX ([3H]RTX) and inhibition of capsaicin-induced Ca2+ influx. α-Spinasterol exhibited an antinociceptive effect to noxious heat, but the mechanical threshold was unaffected. The specific action involving TRPV1 was confirmed by the lack of antinociceptive effect in mice systemically treated with RTX, which is known to ablate TRPV1 expressing neurons. Its effectiveness was demonstrated by its ability to reduce inflammatory hypersensitivity induced by complete Freund’s adjuvant (CFA). Interestingly, the body temperature was unaffected by its antagonistic action. (127)

TRPV3

TRPV3 is a thermosensitive channel that has a high sequence homology with TRPV1. TRPV3 was initially found to be exclusively expressed in the keratinocytes; however, further studies have shown its expression in sensory and central neurons, nasal mucosa, tongue, kidney, and testis. TRPV3 expression has been shown to be enhanced in painful conditions and is being pursued as a target for developing analgesics. (129-135)
Nerve terminals of the sensory neurons in the periphery are surrounded by keratinocytes; therefore substances such as ATP, prostaglandins, and nerve growth factor released from keratinocytes can make the nerve terminals more excitable. (136, 137) TRPV3 knockout animals lacked ATP release from peripheral terminals. (138) Activation of PLC has been shown to modulate the function of TRPV3 by PIP2-mediated mechanism and by IP3-mediated increase in intracellular Ca2+ levels. (138, 139) Overexpression of TRPV3 in animals results in a “hairless” phenotype, indicating the involvement in functions other than nociception. (140, 141)
Several of the TRPV1 activating phytochemicals also activate TRPV3. This may be due to the possibility of coassembly of channels withTRPV1 and TRPV3 subunits. These compounds include thymol, carvacrol, camphor, and eugenol. The channel is also activated by menthol and moderate heat (between 30 and 35 °C). (70, 115, 123, 142) Frankincense is a resin obtained from the frankincense tree (Boswellia thurifera), which belongs to the family Buseraceae. The active ingredient is boswellia acid. Myrrh is another resin from the species commiphora. It has been shown that the active ingredient in these resins, incensole, has psychoactive properties. Incensole smoke is used in religious ceremonies to attain higher levels of meditation. Incensole is an activator of TRPV3. (143)
Compounds in Cannabis sativa, CBD and tetrahydrocannabivarin, caused TRPV3-mediated Ca2+ influx with a EC50 of ∼3.7 μM. Cannabigerovarin and cannabigerolic acid interacted with the channel by causing reduced carvacrol induced Ca2+ influx. (144)

TRPV4

TRPV4 is expressed in hypothalamus, sensory neurons, trachea, kidney, cochlear hair cells, vascular smooth muscle cells, endothelial cells, and keratinocytes. (145-148) TRPV4 is activated by cell-swelling induced by hypotonicity, shear stress, heat (>27 °C), diacyl glycerol (DAG), phorbol esters, 5′,6′-epoxyeicosatrienoic acid (5′,6′-EET), and 4-α-phorbol 12,13-didecanoate (4-α-PDD). (146-151) TRPV4 mediates mechanical sensitivity by direct activation of the channel as well as by second messengers produced by mechanical stimuli. (152-154) The role of TRPV4 in nociception is confirmed by the administration of antisense oligodeoxynucleotide. (155-158)
As discussed under activators of TRPA1, paclitaxel is obtained from the Pacific yew and is used to treat certain forms of cancer (Figure 2). The tactile and cold allodynia induced by paclitaxel have been attributed to its ability to activate both TRPV4 and TRPA1. When antagonists of these TRP channels were administered individually, tactile allodynia induced by paclitaxel was alleviated partially. However, a combination of both completely alleviated tactile allodynia. Paclitaxel-induced CGRP release from mouse esophagus was abolished by TRPA1 and TRPV4 antagonists, suggesting that TRPA1 and TRPV4 contribute to paclitaxel-induced neuropathy. (41) Bisandrographolide and andrographolide are diterpenoids purified from Andrographis paniculata, which belongs to the family Acanthaceae (Figure 2). Bisandrographolide is able to selectively activate TRPV4 without having any effects on TRPV1, TRPV2, and TRPV3. The abundant phytochemical in the extract, andrographolide, failed to activate TRPV4. (159)
Cannabidivarin and tetrahydrocannabivarin from cannabis induced TRPV4-mediated Ca2+ influx with an EC50 of 0.9–6.4 μm. whereas cannabigerolic acid, cannabigerovarin, cannabinol, and cannabigerol interacted with TRPV4 causing reduced 4-α-PDD induced responses. (144)

Concluding Remarks

ARTICLE SECTIONS
Jump To

The TRP family of ion channels has emerged as targets for phytochemicals in botanicals. From the scientific studies, it is becoming apparent that specific and potent active ingredients are being isolated and identified from botanicals. More interestingly, specific receptors for the active ingredients are also being identified, cloned, and characterized. Further, there are endogenous ligands for some of these receptors. The classic examples are endorphin and enkephalin for opioid receptors and anandamide for cannabinoid and TRPV1 receptors. The presence of endogenous ligands strengthens the argument that phytochemicals can fulfill the deficiency of the endogenous ligands or can overactivate the receptor to exert unphysiological responses. TRP channels are expressed in neuronal and nonneuronal cells. Activation of these receptors at the nerve terminals can initiate an afferent sensory signal by depolarizing the nerve terminal and generating an action potential. On the other hand, activation of TRP channels at the nerve terminals can cause an efferent function of releasing peptide hormones that can act locally on other cells in a paracrine fashion (local release) and stimulate the cells to release hormones in an endocrine fashion (release into the blood). Expression of TRP channels in nonneuronal cells, such as pancreatic beta cells and enteroendocrine cells, can release insulin and GLP-1, respectively, which can play a role in glucose homeostasis. It is important to consider the specificity of action, potency, and the bioavailability of these phytochemicals. It is certain that more phytochemicals and their TRP channel targets will be identified in the future to attribute scientific bases for the physiological effects and the health benefits produced by botanicals.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
    • Louis S. Premkumar - Department of Pharmacology, Southern Illinois University School of Medicine, Springfield, Illinois 62702, United States Email: [email protected]
    • Funding

      This work is supported by a grant from NIDA (DA028017).

    • Notes
      The authors declare no competing financial interest.

    Acknowledgment

    ARTICLE SECTIONS
    Jump To

    I thank Somaja Louis for critically reading the manuscript.

    References

    ARTICLE SECTIONS
    Jump To

    This article references 159 other publications.

    1. 1
      Caterina, M. J., Schumacher, M. A., Tominaga, M., Rosen, T. A., Levine, J. D., and Julius, D. (1997) The capsaicin receptor: a heat-activated ion channel in the pain pathway Nature 389, 816 824
    2. 2
      Cosens, D. J. and Manning, A. (1969) Abnormal electroretinogram from a Drosophila mutant Nature 224, 285 287
    3. 3
      Montell, C. and Rubin, G. M. (1989) Molecular characterization of the Drosophila trp locus: a putative integral membrane protein required for phototransduction Neuron 2, 1313 1323
    4. 4
      Montell, C. (2005) The TRP superfamily of cation channels Sci. STKE re3
    5. 5
      Story, G. M., Peier, A. M., Reeve, A. J., Eid, S. R., Mosbacher, J., Hricik, T. R., Earley, T. J., Hergarden, A. C., Andersson, D. A., Hwang, S. W., McIntyre, P., Jegla, T., Bevan, S., and Patapoutian, A. (2003) ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures Cell 112, 819 829
    6. 6
      Jordt, S. E., Bautista, D. M., Chuang, H. H., McKemy, D. D., Zygmunt, P. M., Hogestatt, E. D., Meng, I. D., and Julius, D. (2004) Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1 Nature 427, 260 265
    7. 7
      Peier, A. M., Moqrich, A., Hergarden, A. C., Reeve, A. J., Andersson, D. A., Story, G. M., Earley, T. J., Dragoni, I., McIntyre, P., Bevan, S., and Patapoutian, A. (2002) A TRP channel that senses cold stimuli and menthol Cell 108, 705 715
    8. 8
      McKemy, D. D., Neuhausser, W. M., and Julius, D. (2002) Identification of a cold receptor reveals a general role for TRP channels in thermosensation Nature 416, 52 58
    9. 9
      Cao, E., Liao, M., Cheng, Y., and Julius, D. (2013) TRPV1 structures in distinct conformations reveal activation mechanisms Nature 504, 113 118
    10. 10
      Liao, M., Cao, E., Julius, D., and Cheng, Y. (2013) Structure of the TRPV1 ion channel determined by electron cryo-microscopy Nature 504, 107 112
    11. 11
      Bikman, B. T., Zheng, D., Pories, W. J., Chapman, W., Pender, J. R., Bowden, R. C., Reed, M. A., Cortright, R. N., Tapscott, E. B., Houmard, J. A., Tanner, C. J., Lee, J., and Dohm, G. L. (2008) Mechanism for improved insulin sensitivity after gastric bypass surgery J. Clin. Endocrinol. Metab. 93, 4656 4663
    12. 12
      Corey, D. P., Garcia-Anoveros, J., Holt, J. R., Kwan, K. Y., Lin, S. Y., Vollrath, M. A., Amalfitano, A., Cheung, E. L., Derfler, B. H., Duggan, A., Geleoc, G. S., Gray, P. A., Hoffman, M. P., Rehm, H. L., Tamasauskas, D., and Zhang, D. S. (2004) TRPA1 is a candidate for the mechanosensitive transduction channel of vertebrate hair cells Nature 432, 723 730
    13. 13
      Bandell, M., Story, G. M., Hwang, S. W., Viswanath, V., Eid, S. R., Petrus, M. J., Earley, T. J., Patapoutian, A., Bautista, D. M., Movahed, P., Hinman, A., Axelsson, H. E., Sterner, O., Hogestatt, E. D., Julius, D., Jordt, S. E., and Zygmunt, P. M. (2004) Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin Neuron 41, 849 857
    14. 14
      Bautista, D. M., Movahed, P., Hinman, A., Axelsson, H. E., Sterner, O., Hogestatt, E. D., Julius, D., Jordt, S. E., and Zygmunt, P. M. (2005) Pungent products from garlic activate the sensory ion channel TRPA1 Proc. Natl. Acad. Sci. U.S.A. 102, 12248 12252
    15. 15
      Macpherson, L. J., Geierstanger, B. H., Viswanath, V., Bandell, M., Eid, S. R., Hwang, S., and Patapoutian, A. (2005) The pungency of garlic: activation of TRPA1 and TRPV1 in response to allicin Curr. Biol. 15, 929 934
    16. 16
      Kwan, K. Y., Allchorne, A. J., Vollrath, M. A., Christensen, A. P., Zhang, D. S., Woolf, C. J., and Corey, D. P. (2006) TRPA1 contributes to cold, mechanical, and chemical nociception but is not essential for hair-cell transduction Neuron 50, 277 289
    17. 17
      Andersson, D. A., Gentry, C., Moss, S., and Bevan, S. (2008) Transient receptor potential A1 is a sensory receptor for multiple products of oxidative stress J. Neurosci. 28, 2485 2494
    18. 18
      Bessac, B. F., Sivula, M., von Hehn, C. A., Escalera, J., Cohn, L., and Jordt, S. E. (2008) TRPA1 is a major oxidant sensor in murine airway sensory neurons J. Clin. Invest. 118, 1899 1910
    19. 19
      Cao, D. S., Zhong, L., Hsieh, T. H., Abooj, M., Bishnoi, M., Hughes, L., and Premkumar, L. S. (2012) Expression of transient receptor potential ankyrin 1 (TRPA1) and its role in insulin release from rat pancreatic beta cells PloS One 7, e38005
    20. 20
      Brownlee, M. (2001) Biochemistry and molecular cell biology of diabetic complications Nature 414, 813 820
    21. 21
      Nagata, K., Duggan, A., Kumar, G., and Garcia-Anoveros, J. (2005) Nociceptor and hair cell transducer properties of TRPA1, a channel for pain and hearing J. Neurosci. 25, 4052 4061
    22. 22
      Obata, K., Katsura, H., Mizushima, T., Yamanaka, H., Kobayashi, K., Dai, Y., Fukuoka, T., Tokunaga, A., Tominaga, M., and Noguchi, K. (2005) TRPA1 induced in sensory neurons contributes to cold hyperalgesia after inflammation and nerve injury J. Clin. Invest. 115, 2393 2401
    23. 23
      Hinman, A., Chuang, H. H., Bautista, D. M., and Julius, D. (2006) TRP channel activation by reversible covalent modification Proc. Natl. Acad. Sci. U.S.A. 103, 19564 19568
    24. 24
      Raisinghani, M., Zhong, L., Jeffry, J. A., Bishnoi, M., Pabbidi, R. M., Pimentel, F., Cao, D. S., Evans, M. S., and Premkumar, L. S. (2011) Activation characteristics of transient receptor potential ankyrin 1 and its role in nociception Am. J. Physiol.: Cell Physiol. 301, C587 600
    25. 25
      Premkumar, L. S. and Raisinghani, M. (2006) Nociceptors in cardiovascular functions: complex interplay as a result of cyclooxygenase inhibition Mol. Pain 2, 26
    26. 26
      Khan, A., Safdar, M., Ali Khan, M. M., Khattak, K. N., and Anderson, R. A. (2003) Cinnamon improves glucose and lipids of people with type 2 diabetes Diabetes Care 26, 3215 3218
    27. 27
      Hlebowicz, J., Darwiche, G., Bjorgell, O., and Almer, L. O. (2007) Effect of cinnamon on postprandial blood glucose, gastric emptying, and satiety in healthy subjects Am. J. Clin. Nutr. 85, 1552 1556
    28. 28
      Lissiman, E., Bhasale, A. L., and Cohen, M. (2009) Garlic for the common cold Cochrane Database Syst. Rev. 3, CD006206
    29. 29
      Lissiman, E., Bhasale, A. L., and Cohen, M. (2012) Garlic for the common cold Cochrane Database Syst. Rev. 14, CD006206
    30. 30
      Cutler, R. R., Odent, M., Hajj-Ahmad, H., Maharjan, S., Bennett, N. J., Josling, P. D., Ball, V., Hatton, P., and Dall’Antonia, M. (2009) In vitro activity of an aqueous allicin extract and a novel allicin topical gel formulation against Lancefield group B streptococci J. Antimicrob. Chemother. 63, 151 154
    31. 31
      Khanum, F., Anilakumar, K. R., and Viswanathan, K. R. (2004) Anticarcinogenic properties of garlic: a review Crit. Rev. Food. Sci. Nutr. 44, 479 488
    32. 32
      Shenoy, N. R. and Choughuley, A. S. (1992) Inhibitory effect of diet related sulphydryl compounds on the formation of carcinogenic nitrosamines Cancer Lett. 65, 227 232
    33. 33
      Koizumi, K., Iwasaki, Y., Narukawa, M., Iitsuka, Y., Fukao, T., Seki, T., Ariga, T., and Watanabe, T. (2009) Diallyl sulfides in garlic activate both TRPA1 and TRPV1 Biochem. Biophys. Res. Commun. 382, 545 548
    34. 34
      Leamy, A. W., Shukla, P., McAlexander, M. A., Carr, M. J., and Ghatta, S. (2011) Curcumin ((E,E)-1,7-bis(4-hydroxy-3-methoxyphenyl)-1,6-heptadiene-3,5-dione) activates and desensitizes the nociceptor ion channel TRPA1 Neurosci. Lett. 503, 157 162
    35. 35
      Gupta, S. C., Patchva, S., and Aggarwal, B. B. (2013) Therapeutic roles of curcumin: lessons learned from clinical trials AAPS J. 15, 195 218
    36. 36
      Song, Y., Sonawane, N. D., Salinas, D., Qian, L., Pedemonte, N., Galietta, L. J., and Verkman, A. S. (2004) Evidence against the rescue of defective DeltaF508-CFTR cellular processing by curcumin in cell culture and mouse models J. Biol. Chem. 279, 40629 40633
    37. 37
      Gao, J., Zhou, H., Lei, T., Zhou, L., Li, W., Li, X., and Yang, B. (2011) Curcumin inhibits renal cyst formation and enlargement in vitro by regulating intracellular signaling pathways Eur. J. Pharmacol. 654, 92 99
    38. 38
      Sohma, Y., Yu, Y. C., and Hwang, T. C. (2013) Curcumin and genistein: the combined effects on disease-associated CFTR mutants and their clinical implications Curr. Pharm. Des. 19, 3521 3528
    39. 39
      Nassini, R., Materazzi, S., Vriens, J., Prenen, J., Benemei, S., De Siena, G., la Marca, G., Andre, E., Preti, D., Avonto, C., Sadofsky, L., Di Marzo, V., De Petrocellis, L., Dussor, G., Porreca, F., Taglialatela-Scafati, O., Appendino, G., Nilius, B., and Geppetti, P. (2012) The ‘headache tree’ via umbellulone and TRPA1 activates the trigeminovascular system Brain 135, 376 390
    40. 40
      Zhong, J., Pollastro, F., Prenen, J., Zhu, Z., Appendino, G., and Nilius, B. (2011) Ligustilide: a novel TRPA1 modulator Pfluegers Arch. 462, 841 849
    41. 41
      Materazzi, S., Fusi, C., Benemei, S., Pedretti, P., Patacchini, R., Nilius, B., Prenen, J., Creminon, C., Geppetti, P., and Nassini, R. (2012) TRPA1 and TRPV4 mediate paclitaxel-induced peripheral neuropathy in mice via a glutathione-sensitive mechanism Pfluegers Arch. 463, 561 569
    42. 42
      De Petrocellis, L., Vellani, V., Schiano-Moriello, A., Marini, P., Magherini, P. C., Orlando, P., and Di Marzo, V. (2008) Plant-derived cannabinoids modulate the activity of transient receptor potential channels of ankyrin type-1 and melastatin type-8 J. Pharmacol. Exp. Ther. 325, 1007 1015
    43. 43
      Montrucchio, D. P., Cordova, M. M., and Santos, A. R. (2013) Plant derived aporphinic alkaloid S-(+)-dicentrine induces antinociceptive effect in both acute and chronic inflammatory pain models: evidence for a role of TRPA1 channels PloS One 8, e67730
    44. 44
      Talavera, K., Gees, M., Karashima, Y., Meseguer, V. M., Vanoirbeek, J. A., Damann, N., Everaerts, W., Benoit, M., Janssens, A., Vennekens, R., Viana, F., Nemery, B., Nilius, B., and Voets, T. (2009) Nicotine activates the chemosensory cation channel TRPA1 Nat. Neurosci. 12, 1293 1299
    45. 45
      Riera, C. E., Menozzi-Smarrito, C., Affolter, M., Michlig, S., Munari, C., Robert, F., Vogel, H., Simon, S. A., and Le Coutre, J. (2009) Compounds from Sichuan and Melegueta peppers activate, covalently and non-covalently, TRPA1 and TRPV1 channels Br. J. Pharmacol. 157, 1398 1409
    46. 46
      Vazquez, G., Wedel, B. J., Aziz, O., Trebak, M., and Putney, J. W., Jr. (2004) The mammalian TRPC cation channels Biochim. Biophys. Acta 1742, 21 36
    47. 47
      Birnbaumer, L. (2009) The TRPC class of ion channels: a critical review of their roles in slow, sustained increases in intracellular Ca(2+) concentrations Annu. Rev. Pharmacol. Toxicol. 49, 395 426
    48. 48
      Leuner, K., Kazanski, V., Muller, M., Essin, K., Henke, B., Gollasch, M., Harteneck, C., and Muller, W. E. (2007) Hyperforin--a key constituent of St. John’s wort specifically activates TRPC6 channels FASEB J. 21, 4101 4111
    49. 49
      Sikand, P. and Premkumar, L. S. (2007) Potentiation of glutamatergic synaptic transmission by protein kinase C-mediated sensitization of TRPV1 at the first sensory synapse J. Physiol. 581, 631 647
    50. 50
      Cao, D. S., Yu, S. Q., and Premkumar, L. S. (2009) Modulation of transient receptor potential Vanilloid 4-mediated membrane currents and synaptic transmission by protein kinase C Mol. Pain 5, 5
    51. 51
      Jeffry, J. A., Yu, S. Q., Sikand, P., Parihar, A., Evans, M. S., and Premkumar, L. S. (2009) Selective targeting of TRPV1 expressing sensory nerve terminals in the spinal cord for long lasting analgesia PloS One 4, e7021
    52. 52
      Evans, M. S., Cheng, X., Jeffry, J. A., Disney, K. E., and Premkumar, L. S. (2012) Sumatriptan inhibits TRPV1 channels in trigeminal neurons Headache 52, 773 784
    53. 53
      Zhou, J., Du, W., Zhou, K., Tai, Y., Yao, H., Jia, Y., and Ding, Y. (2008) Critical role of TRPC6 channels in the formation of excitatory synapses Nat. Neurosci. 11, 741 743
    54. 54
      Nilius, B. and Owsianik, G. (2011) The transient receptor potential family of ion channels Genome Biol. 12, 218
    55. 55
      Carakostas, M. C., Curry, L. L., Boileau, A. C., and Brusick, D. J. (2008) Overview: the history, technical function and safety of rebaudioside A, a naturally occurring steviol glycoside, for use in food and beverages Food Chem. Toxicol. 46 (Suppl 7) S1 S10
    56. 56
      Medler, K. F. (2011) Multiple roles for TRPs in the taste system: not your typical TRPs Adv. Exp. Med. Biol. 704, 831 846
    57. 57
      Sprous, D. and Palmer, K. R. (2010) The T1R2/T1R3 sweet receptor and TRPM5 ion channel taste targets with therapeutic potential Prog. Mol. Biol. Transl. Sci. 91, 151 208
    58. 58
      Palmer, R. K. (2007) The pharmacology and signaling of bitter, sweet, and umami taste sensing Mol. Interv. 7, 87 98
    59. 59
      Depoortere, I. (2014) Taste receptors of the gut: emerging roles in health and disease Gut 63, 179 190
    60. 60
      Kaji, I., Karaki, S. I., and Kuwahara, A. (2014) Taste Sensing in the Colon Curr. Pharm. Des. 20, 2766 2774
    61. 61
      Premkumar, L. S., Raisinghani, M., Pingle, S. C., Long, C., and Pimentel, F. (2005) Downregulation of transient receptor potential melastatin 8 by protein kinase C-mediated dephosphorylation J. Neurosci. 25, 11322 11329
    62. 62
      Tsuzuki, K., Xing, H., Ling, J., and Gu, J. G. (2004) Menthol-induced Ca2+ release from presynaptic Ca2+ stores potentiates sensory synaptic transmission J. Neurosci. 24, 762 771
    63. 63
      Tsavaler, L., Shapero, M. H., Morkowski, S., and Laus, R. (2001) Trp-p8, a novel prostate-specific gene, is up-regulated in prostate cancer and other malignancies and shares high homology with transient receptor potential calcium channel proteins Cancer Res. 61, 3760 3769
    64. 64
      Uhl, G. R., Walther, D., Behm, F. M., and Rose, J. E. (2011) Menthol preference among smokers: association with TRPA1 variants Nicotine Tob. Res. 13, 1311 1315
    65. 65
      Hans, M., Wilhelm, M., and Swandulla, D. (2012) Menthol suppresses nicotinic acetylcholine receptor functioning in sensory neurons via allosteric modulation Chem. Senses 37, 463 469
    66. 66
      Willis, D. N., Liu, B., Ha, M. A., Jordt, S. E., and Morris, J. B. (2011) Menthol attenuates respiratory irritation responses to multiple cigarette smoke irritants FASEB J. 25, 4434 4444
    67. 67
      Leuenroth, S. J., Okuhara, D., Shotwell, J. D., Markowitz, G. S., Yu, Z., Somlo, S., and Crews, C. M. (2007) Triptolide is a traditional Chinese medicine-derived inhibitor of polycystic kidney disease Proc. Natl. Acad. Sci. U.S.A. 104, 4389 4394
    68. 68
      Szallasi, A. and Blumberg, P. M. (1999) Vanilloid (Capsaicin) receptors and mechanisms Pharmacol. Rev. 51, 159 212
    69. 69
      Vennekens, R., Owsianik, G., and Nilius, B. (2008) Vanilloid transient receptor potential cation channels: an overview Curr. Pharm. Des. 14, 18 31
    70. 70
      Vriens, J., Nilius, B., and Vennekens, R. (2008) Herbal compounds and toxins modulating TRP channels Curr. Neuropharmacol. 6, 79 96
    71. 71
      Hellwig, N., Albrecht, N., Harteneck, C., Schultz, G., and Schaefer, M. (2005) Homo- and heteromeric assembly of TRPV channel subunits J. Cell Sci. 118, 917 928
    72. 72
      Tominaga, M., Caterina, M. J., Malmberg, A. B., Rosen, T. A., Gilbert, H., Skinner, K., Raumann, B. E., Basbaum, A. I., and Julius, D. (1998) The cloned capsaicin receptor integrates multiple pain-producing stimuli Neuron 21, 531 543
    73. 73
      Nakatsuka, T., Furue, H., Yoshimura, M., and Gu, J. G. (2002) Activation of central terminal vanilloid receptor-1 receptors and alpha beta-methylene-ATP-sensitive P2X receptors reveals a converged synaptic activity onto the deep dorsal horn neurons of the spinal cord J. Neurosci. 22, 1228 1237
    74. 74
      Baccei, M. L., Bardoni, R., and Fitzgerald, M. (2003) Development of nociceptive synaptic inputs to the neonatal rat dorsal horn: glutamate release by capsaicin and menthol J. Physiol. 549, 231 242
    75. 75
      Kim, Y. H., Back, S. K., Davies, A. J., Jeong, H., Jo, H. J., Chung, G., Na, H. S., Bae, Y. C., Kim, S. J., Kim, J. S., Jung, S. J., and Oh, S. B. (2012) TRPV1 in GABAergic interneurons mediates neuropathic mechanical allodynia and disinhibition of the nociceptive circuitry in the spinal cord Neuron 74, 640 647
    76. 76
      Doyle, M. W., Bailey, T. W., Jin, Y. H., and Andresen, M. C. (2002) Vanilloid receptors presynaptically modulate cranial visceral afferent synaptic transmission in nucleus tractus solitarius J. Neurosci. 22, 8222 8229
    77. 77
      Marinelli, S., Di Marzo, V., Berretta, N., Matias, I., Maccarrone, M., Bernardi, G., and Mercuri, N. B. (2003) Presynaptic facilitation of glutamatergic synapses to dopaminergic neurons of the rat substantia nigra by endogenous stimulation of vanilloid receptors J. Neurosci. 23, 3136 3144
    78. 78
      Marinelli, S., Vaughan, C. W., Christie, M. J., and Connor, M. (2002) Capsaicin activation of glutamatergic synaptic transmission in the rat locus coeruleus in vitro J. Physiol. 543, 531 540
    79. 79
      Gibson, H. E., Edwards, J. G., Page, R. S., Van Hook, M. J., and Kauer, J. A. (2008) TRPV1 channels mediate long-term depression at synapses on hippocampal interneurons Neuron 57, 746 759
    80. 80
      Grueter, B. A., Brasnjo, G., and Malenka, R. C. (2010) Postsynaptic TRPV1 triggers cell type-specific long-term depression in the nucleus accumbens Nat. Neurosci. 13, 1519 1525
    81. 81
      Premkumar, L. S. and Ahern, G. P. (2000) Induction of vanilloid receptor channel activity by protein kinase C Nature 408, 985 990
    82. 82
      Caterina, M. J. and Julius, D. (2001) The vanilloid receptor: a molecular gateway to the pain pathway Annu. Rev. Neurosci. 24, 487 517
    83. 83
      Minke, B. and Cook, B. (2002) TRP channel proteins and signal transduction Physiol. Rev. 82, 429 472
    84. 84
      Clapham, D. E. (2003) TRP channels as cellular sensors Nature 426, 517 524
    85. 85
      Premkumar, L. S. and Bishnoi, M. (2011) Disease-related changes in TRPV1 expression and its implications for drug development Curr. Top. Med. Chem. 11, 2192 2209
    86. 86
      Julius, D. (2013) TRP channels and pain Annu. Rev. Cell Dev. Biol. 29, 355 384
    87. 87
      Mezey, E., Toth, Z. E., Cortright, D. N., Arzubi, M. K., Krause, J. E., Elde, R., Guo, A., Blumberg, P. M., and Szallasi, A. (2000) Distribution of mRNA for vanilloid receptor subtype 1 (VR1), and VR1-like immunoreactivity, in the central nervous system of the rat and human Proc. Natl. Acad. Sci. U.S.A. 97, 3655 3660
    88. 88
      Huang, S. M., Bisogno, T., Trevisani, M., Al-Hayani, A., De Petrocellis, L., Fezza, F., Tognetto, M., Petros, T. J., Krey, J. F., Chu, C. J., Miller, J. D., Davies, S. N., Geppetti, P., Walker, J. M., and Di Marzo, V. (2002) An endogenous capsaicin-like substance with high potency at recombinant and native vanilloid VR1 receptors Proc. Natl. Acad. Sci. U.S.A. 99, 8400 8405
    89. 89
      Mishra, S. K., Tisel, S. M., Orestes, P., Bhangoo, S. K., and Hoon, M. A. (2011) TRPV1-lineage neurons are required for thermal sensation EMBO J. 30, 582 593
    90. 90
      Cavanaugh, D. J., Chesler, A. T., Braz, J. M., Shah, N. M., Julius, D., and Basbaum, A. I. (2011) Restriction of transient receptor potential vanilloid-1 to the peptidergic subset of primary afferent neurons follows its developmental downregulation in nonpeptidergic neurons J. Neurosci. 31, 10119 10127
    91. 91
      Lundberg, J. M., Martling, C. R., and Saria, A. (1983) Substance P and capsaicin-induced contraction of human bronchi Acta Physiol. Scand. 119, 49 53
    92. 92
      Mitchell, J. A., Williams, F. M., Williams, T. J., and Larkin, S. W. (1997) Role of nitric oxide in the dilator actions of capsaicin-sensitive nerves in the rabbit coronary circulation Neuropeptides 31, 333 338
    93. 93
      Zygmunt, P. M., Petersson, J., Andersson, D. A., Chuang, H., Sorgard, M., Di Marzo, V., Julius, D., Hogestatt, E. D., and Wang, Y. (1999) Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide Nature 400, 452 457
    94. 94
      Birder, L. A., Nakamura, Y., Kiss, S., Nealen, M. L., Barrick, S., Kanai, A. J., Wang, E., Ruiz, G., De Groat, W. C., Apodaca, G., Watkins, S., and Caterina, M. J. (2002) Altered urinary bladder function in mice lacking the vanilloid receptor TRPV1 Nat. Neurosci. 5, 856 860
    95. 95
      Cruz, F. and Dinis, P. (2007) Resiniferatoxin and botulinum toxin type A for treatment of lower urinary tract symptoms Neurourol. Urodyn. 26, 920 927
    96. 96
      Varga, A., Nemeth, J., Szabo, A., McDougall, J. J., Zhang, C., Elekes, K., Pinter, E., Szolcsanyi, J., and Helyes, Z. (2005) Effects of the novel TRPV1 receptor antagonist SB366791 in vitro and in vivo in the rat Neurosci. Lett. 385, 137 142
    97. 97
      Gavva, N. R. (2008) Body-temperature maintenance as the predominant function of the vanilloid receptor TRPV1 Trends Pharmacol. Sci. 29, 550 557
    98. 98
      Lehto, S. G., Tamir, R., Deng, H., Klionsky, L., Kuang, R., Le, A., Lee, D., Louis, J. C., Magal, E., Manning, B. H., Rubino, J., Surapaneni, S., Tamayo, N., Wang, T., Wang, J., Wang, J., Wang, W., Youngblood, B., Zhang, M., Zhu, D., Norman, M. H., and Gavva, N. R. (2008) Antihyperalgesic effects of (R,E)-N-(2-hydroxy-2,3-dihydro-1H-inden-4-yl)-3-(2-(piperidin-1-yl)-4-(trifluorom ethyl)phenyl)-acrylamide (AMG8562), a novel transient receptor potential vanilloid type 1 modulator that does not cause hyperthermia in rats J. Pharmacol. Exp. Ther. 326, 218 229
    99. 99
      Pabbidi, R. M., Yu, S. Q., Peng, S., Khardori, R., Pauza, M. E., and Premkumar, L. S. (2008) Influence of TRPV1 on diabetes-induced alterations in thermal pain sensitivity Mol. Pain 4, 9
    100. 100
      Van Buren, J. J., Bhat, S., Rotello, R., Pauza, M. E., and Premkumar, L. S. (2005) Sensitization and translocation of TRPV1 by insulin and IGF-I Mol. Pain 1, 17
    101. 101
      Bishnoi, M., Bosgraaf, C. A., Abooj, M., Zhong, L., and Premkumar, L. S. (2011) Streptozotocin-induced early thermal hyperalgesia is independent of glycemic state of rats: role of transient receptor potential vanilloid 1(TRPV1) and inflammatory mediators Mol. Pain 7, 52
    102. 102
      Akiba, Y., Kato, S., Katsube, K., Nakamura, M., Takeuchi, K., Ishii, H., and Hibi, T. (2004) Transient receptor potential vanilloid subfamily 1 expressed in pancreatic islet beta cells modulates insulin secretion in rats Biochem. Biophys. Res. Commun. 321, 219 225
    103. 103
      Razavi, R., Chan, Y., Afifiyan, F. N., Liu, X. J., Wan, X., Yantha, J., Tsui, H., Tang, L., Tsai, S., Santamaria, P., Driver, J. P., Serreze, D., Salter, M. W., and Dosch, H. M. (2006) TRPV1+ sensory neurons control beta cell stress and islet inflammation in autoimmune diabetes Cell 127, 1123 1135
    104. 104
      Wang, X., Miyares, R. L., and Ahern, G. P. (2005) Oleoylethanolamide excites vagal sensory neurones, induces visceral pain and reduces short-term food intake in mice via capsaicin receptor TRPV1 J. Physiol. 564, 541 547
    105. 105
      Ahern, G. P. (2013) Transient receptor potential channels and energy homeostasis Trends Endocrinol. Metab. 24, 554 560
    106. 106
      Kindt, S., Vos, R., Blondeau, K., and Tack, J. (2009) Influence of intra-oesophageal capsaicin instillation on heartburn induction and oesophageal sensitivity in man Neurogastroenterol. Motil. 21, 1032 e1082
    107. 107
      Okumi, H., Tashima, K., Matsumoto, K., Namiki, T., Terasawa, K., and Horie, S. (2012) Dietary agonists of TRPV1 inhibit gastric acid secretion in mice Planta Med. 78, 1801 1806
    108. 108
      Raisinghani, M., Pabbidi, R. M., and Premkumar, L. S. (2005) Activation of transient receptor potential vanilloid 1 (TRPV1) by resiniferatoxin J. Physiol. 567, 771 786
    109. 109
      Roberts, J. C., Davis, J. B., and Benham, C. D. (2004) [3H]Resiniferatoxin autoradiography in the CNS of wild-type and TRPV1 null mice defines TRPV1 (VR-1) protein distribution Brain Res. 995, 176 183
    110. 110
      Chou, M. Z., Mtui, T., Gao, Y. D., Kohler, M., and Middleton, R. E. (2004) Resiniferatoxin binds to the capsaicin receptor (TRPV1) near the extracellular side of the S4 transmembrane domain Biochemistry 43, 2501 2511
    111. 111
      Jung, J., Lee, S. Y., Hwang, S. W., Cho, H., Shin, J., Kang, Y. S., Kim, S., and Oh, U. (2002) Agonist recognition sites in the cytosolic tails of vanilloid receptor 1 J. Biol. Chem. 277, 44448 44454
    112. 112
      Gavva, N. R., Klionsky, L., Qu, Y., Shi, L., Tamir, R., Edenson, S., Zhang, T. J., Viswanadhan, V. N., Toth, A., Pearce, L. V., Vanderah, T. W., Porreca, F., Blumberg, P. M., Lile, J., Sun, Y., Wild, K., Louis, J. C., and Treanor, J. J. (2004) Molecular determinants of vanilloid sensitivity in TRPV1 J. Biol. Chem. 279, 20283 20295
    113. 113
      Brown, D. C., Iadarola, M. J., Perkowski, S. Z., Erin, H., Shofer, F., Laszlo, K. J., Olah, Z., and Mannes, A. J. (2005) Physiologic and antinociceptive effects of intrathecal resiniferatoxin in a canine bone cancer model Anesthesiology 103, 1052 1059
    114. 114
      Iadarola, M. J. and Mannes, A. J. (2011) The vanilloid agonist resiniferatoxin for interventional-based pain control Curr. Top. Med. Chem. 11, 2171 2179
    115. 115
      Yang, B. H., Piao, Z. G., Kim, Y. B., Lee, C. H., Lee, J. K., Park, K., Kim, J. S., and Oh, S. B. (2003) (2003) Activation of vanilloid receptor 1 (VR1) by eugenol J. Dent. Res. 82, 781 785
    116. 116
      Behrendt, H. J., Germann, T., Gillen, C., Hatt, H., and Jostock, R. (2004) Characterization of the mouse cold-menthol receptor TRPM8 and vanilloid receptor type-1 VR1 using a fluorometric imaging plate reader (FLIPR) assay Br. J. Pharmacol. 141, 737 45
    117. 117
      McNamara, F. N., Randall, A., and Gunthorpe, M. J. (2005) Effects of piperine, the pungent component of black pepper, at the human vanilloid receptor (TRPV1) Br. J. Pharmacol. 144, 781 790
    118. 118
      Bhardwaj, R. K., Glaeser, H., Becquemont, L., Klotz, U., Gupta, S. K., and Fromm, M. F. (2002) Piperine, a major constituent of black pepper, inhibits human P-glycoprotein and CYP3A4 J. Pharmacol. Exp. Ther. 302, 645 650
    119. 119
      Bisogno, T., Hanus, L., De Petrocellis, L., Tchilibon, S., Ponde, D. E., Brandi, I., Moriello, A. S., Davis, J. B., Mechoulam, R., and Di Marzo, V. (2001) Molecular targets for cannabidiol and its synthetic analogues: effect on vanilloid VR1 receptors and on the cellular uptake and enzymatic hydrolysis of anandamide Br. J. Pharmacol. 134, 845 852
    120. 120
      Xu, H., Blair, N. T., and Clapham, D. E. (2005) Camphor activates and strongly desensitizes the transient receptor potential vanilloid subtype 1 channel in a vanilloid-independent mechanism J. Neurosci. 25, 8924 8937
    121. 121
      Pearce, L. V., Petukhov, P. A., Szabo, T., Kedei, N., Bizik, F., Kozikowski, A. P., and Blumberg, P. M. (2004) Evodiamine functions as an agonist for the vanilloid receptor TRPV1 Org. Biomol. Chem. 2, 2281 2286
    122. 122
      Morera, E., De Petrocellis, L., Morera, L., Moriello, A. S., Nalli, M., Di Marzo, V., and Ortar, G. (2012) Synthesis and biological evaluation of [6]-gingerol analogues as transient receptor potential channel TRPV1 and TRPA1 modulators Bioorg. Med. Chem. Lett. 22, 1674 1677
    123. 123
      Xu, H., Delling, M., Jun, J. C., and Clapham, D. E. (2006) Oregano, thyme and clove-derived flavors and skin sensitizers activate specific TRP channels Nat. Neurosci. 9, 628 635
    124. 124
      Andre, E., Campi, B., Trevisani, M., Ferreira, J., Malheiros, A., Yunes, R. A., Calixto, J. B., and Geppetti, P. (2006) Pharmacological characterisation of the plant sesquiterpenes polygodial and drimanial as vanilloid receptor agonists Biochem. Pharmacol. 71, 1248 1254
    125. 125
      Iwasaki, Y., Tanabe, M., Kayama, Y., Abe, M., Kashio, M., Koizumi, K., Okumura, Y., Morimitsu, Y., Tominaga, M., Ozawa, Y., and Watanabe, T. (2009) Miogadial and miogatrial with alpha,beta-unsaturated 1,4-dialdehyde moieties—novel and potent TRPA1 agonists Life Sci. 85, 60 69
    126. 126
      Lübbert, M., Kyereme, J., Schöbel, N., Beltrán, L., Wetzel, C. H., and Hatt, H. (2013) Transient receptor potential channels encode volatile chemicals sensed by rat trigeminal ganglion neurons PLoS One 8, e77998
    127. 127
      Trevisan, G., Rossato, M. F., Walker, C. I., Klafke, J. Z., Rosa, F., Oliveira, S. M., Tonello, R., Guerra, G. P., Boligon, A. A., Zanon, R. B., Athayde, M. L., and Ferreira, J. (2012) Identification of the plant steroid alpha-spinasterol as a novel transient receptor potential vanilloid 1 antagonist with antinociceptive properties J. Pharmacol. Exp. Ther. 343, 258 269
    128. 128
      Birnbaumer, L., Yildirim, E., and Abramowitz, J. (2003) A comparison of the genes coding for canonical TRP channels and their M, V and P relatives Cell Calcium 33, 419 432
    129. 129
      Peier, A. M., Reeve, A. J., Andersson, D. A., Moqrich, A., Earley, T. J., Hergarden, A. C., Story, G. M., Colley, S., Hogenesch, J. B., McIntyre, P., Bevan, S., and Patapoutian, A. (2002) A heat-sensitive TRP channel expressed in keratinocytes Science 296, 2046 2049
    130. 130
      Moqrich, A., Hwang, S. W., Earley, T. J., Petrus, M. J., Murray, A. N., Spencer, K. S., Andahazy, M., Story, G. M., and Patapoutian, A. (2005) Impaired thermosensation in mice lacking TRPV3, a heat and camphor sensor in the skin Science 307, 1468 1472
    131. 131
      Xu, H., Ramsey, I. S., Kotecha, S. A., Moran, M. M., Chong, J. A., Lawson, D., Ge, P., Lilly, J., Silos-Santiago, I., Xie, Y., DiStefano, P. S., Curtis, R., and Clapham, D. E. (2002) TRPV3 is a calcium-permeable temperature-sensitive cation channel Nature 418, 181 186
    132. 132
      Smith, G. D., Gunthorpe, M. J., Kelsell, R. E., Hayes, P. D., Reilly, P., Facer, P., Wright, J. E., Jerman, J. C., Walhin, J. P., Ooi, L., Egerton, J., Charles, K. J., Smart, D., Randall, A. D., Anand, P., and Davis, J. B. (2002) TRPV3 is a temperature-sensitive vanilloid receptor-like protein Nature 418, 186 190
    133. 133
      Gopinath, P., Wan, E., Holdcroft, A., Facer, P., Davis, J. B., Smith, G. D., Bountra, C., and Anand, P. (2005) Increased capsaicin receptor TRPV1 in skin nerve fibres and related vanilloid receptors TRPV3 and TRPV4 in keratinocytes in human breast pain BMC Women’s Health 5, 2
    134. 134
      Facer, P., Casula, M. A., Smith, G. D., Benham, C. D., Chessell, I. P., Bountra, C., Sinisi, M., Birch, R., and Anand, P. (2007) Differential expression of the capsaicin receptor TRPV1 and related novel receptors TRPV3, TRPV4 and TRPM8 in normal human tissues and changes in traumatic and diabetic neuropathy BMC Neurol. 7, 11
    135. 135
      Huang, S. M., Lee, H., Chung, M. K., Park, U., Yu, Y. Y., Bradshaw, H. B., Coulombe, P. A., Walker, J. M., and Caterina, M. J. (2008) Overexpressed transient receptor potential vanilloid 3 ion channels in skin keratinocytes modulate pain sensitivity via prostaglandin E2 J. Neurosci. 28, 13727 13737
    136. 136
      Hammarstrom, S., Hamberg, M., Samuelsson, B., Duell, E. A., Stawiski, M., and Voorhees, J. J. (1975) Increased concentrations of nonesterified arachidonic acid, 12L-hydroxy-5,8,10,14-eicosatetraenoic acid, prostaglandin E2, and prostaglandin F2alpha in epidermis of psoriasis Proc. Natl. Acad. Sci. U.S.A. 72, 5130 5134
    137. 137
      Brash, A. R. (2001) Arachidonic acid as a bioactive molecule J. Clin. Invest. 107, 1339 1345
    138. 138
      Mandadi, S., Sokabe, T., Shibasaki, K., Katanosaka, K., Mizuno, A., Moqrich, A., Patapoutian, A., Fukumi-Tominaga, T., Mizumura, K., and Tominaga, M. (2009) TRPV3 in keratinocytes transmits temperature information to sensory neurons via ATP Pfluegers Arch. 458, 1093 2002
    139. 139
      Xiao, R., Tang, J., Wang, C., Colton, C. K., Tian, J., and Zhu, M. X. (2008) Calcium plays a central role in the sensitization of TRPV3 channel to repetitive stimulations J. Biol. Chem. 283, 6162 6174
    140. 140
      Imura, K. (2007) Influence of TRPV3 mutation on hair growth cycle in mice Biochem. Biophys. Res. Commun. 363, 479 483
    141. 141
      Asakawa, M., Yoshioka, T., Matsutani, T., Hikita, I., Suzuki, M., Oshima, I., Tsukahara, K., Arimura, A., Horikawa, T., Hirasawa, T., and Sakata, T. (2006) Association of a mutation in TRPV3 with defective hair growth in rodents J. Invest. Dermatol. 126, 2664 2672
    142. 142
      Macpherson, L. J., Hwang, S. W., Miyamoto, T., Dubin, A. E., Patapoutian, A., and Story, G. M. (2006) More than cool: promiscuous relationships of menthol and other sensory compounds Mol. Cell Neurosci. 32, 335 343
    143. 143
      Moussaieff, A., Rimmerman, N., Bregman, T., Straiker, A., Felder, C. C., Shoham, S., Kashman, Y., Huang, S. M., Lee, H., Shohami, E., Mackie, K., Caterina, M. J., Walker, J. M., Fride, E., and Mechoulam, R. (2008) Incensole acetate, an incense component, elicits psychoactivity by activating TRPV3 channels in the brain FASEB J. 22, 3024 3034
    144. 144
      De Petrocellis, L., Orlando, P., Moriello, A. S., Aviello, G., Stott, C., Izzo, A. A., and Di Marzo, V. (2012) Cannabinoid actions at TRPV channels: effects on TRPV3 and TRPV4 and their potential relevance to gastrointestinal inflammation Acta Physiol. 204, 255 266
    145. 145
      Strotmann, R., Harteneck, C., Nunnenmacher, K., Schultz, G., and Plant, T. D. (2000) OTRPC4, a nonselective cation channel that confers sensitivity to extracellular osmolarity Nat. Cell Biol. 2, 695 702
    146. 146
      Watanabe, H., Davis, J. B., Smart, D., Jerman, J. C., Smith, G. D., Hayes, P., Vriens, J., Cairns, W., Wissenbach, U., Prenen, J., Flockerzi, V., Droogmans, G., Benham, C. D., and Nilius, B. (2002) Activation of TRPV4 channels (hVRL-2/mTRP12) by phorbol derivatives J. Biol. Chem. 277, 13569 13577
    147. 147
      Suzuki, M., Mizuno, A., Kodaira, K., and Imai, M. (2003) Impaired pressure sensation in mice lacking TRPV4 J. Biol. Chem. 278, 22664 22668
    148. 148
      Nilius, B., Droogmans, G., and Wondergem, R. (2003) Transient receptor potential channels in endothelium: solving the calcium entry puzzle? Endothelium 10, 5 15
    149. 149
      Kohler, R., Heyken, W. T., Heinau, P., Schubert, R., Si, H., Kacik, M., Busch, C., Grgic, I., Maier, T., and Hoyer, J. (2006) Evidence for a functional role of endothelial transient receptor potential V4 in shear stress-induced vasodilatation Arterioscler., Thromb., Vasc. Biol. 26, 1495 15502
    150. 150
      Watanabe, H., Vriens, J., Suh, S. H., Benham, C. D., Droogmans, G., and Nilius, B. (2002) Heat-evoked activation of TRPV4 channels in a HEK293 cell expression system and in native mouse aorta endothelial cells J. Biol. Chem. 277, 47044 47051
    151. 151
      Watanabe, H., Vriens, J., Prenen, J., Droogmans, G., Voets, T., and Nilius, B. (2003) Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels Nature 424, 434 438
    152. 152
      Loukin, S. H., Su, Z., and Kung, C. (2009) Hypotonic shocks activate rat TRPV4 in yeast in the absence of polyunsaturated fatty acids FEBS Lett. 583, 754 758
    153. 153
      Cao, D. S., Yu, S. Q., and Premkumar, L. S. (2009) Modulation of transient receptor potential Vanilloid 4-mediated membrane currents and synaptic transmission by protein kinase C Mol. Pain 5, 5
    154. 154
      Matthews, B. D., Thodeti, C. K., Tytell, J. D., Mammoto, A., Overby, D. R., and Ingber, D. E. (2010) Ultra-rapid activation of TRPV4 ion channels by mechanical forces applied to cell surface beta1 integrins Integr. Biol. 2, 435 442
    155. 155
      Alessandri-Haber, N., Dina, O. A., Joseph, E. K., Reichling, D. B., and Levine, J. D. (2008) Interaction of transient receptor potential vanilloid 4, integrin, and SRC tyrosine kinase in mechanical hyperalgesia J. Neurosci. 28, 1046 1057
    156. 156
      Chen, X., Alessandri-Haber, N., and Levine, J. D. (2007) Marked attenuation of inflammatory mediator-induced C-fiber sensitization for mechanical and hypotonic stimuli in TRPV4–/– mice Mol. Pain 3, 31
    157. 157
      Alessandri-Haber, N., Yeh, J. J., Boyd, A. E., Parada, C. A., Chen, X., Reichling, D. B., and Levine, J. D. (2003) Hypotonicity induces TRPV4-mediated nociception in rat Neuron 39, 497 511
    158. 158
      Alessandri-Haber, N., Joseph, E., Dina, O. A., Liedtke, W., and Levine, J. D. (2005) TRPV4 mediates pain-related behavior induced by mild hypertonic stimuli in the presence of inflammatory mediator Pain 118, 70 79
    159. 159
      Smith, P. L., Maloney, K. N., Pothen, R. G., Clardy, J., and Clapham, D. E. (2006) Bisandrographolide from Andrographis paniculata activates TRPV4 channels J. Biol. Chem. 281, 29897 29904

    Cited By

    ARTICLE SECTIONS
    Jump To

    This article is cited by 60 publications.

    1. Shatha K. Alhadyan, Vijay Sivaraman, Rob U. Onyenwoke. E-cigarette Flavors, Sensory Perception, and Evoked Responses. Chemical Research in Toxicology 2022, 35 (12) , 2194-2209. https://doi.org/10.1021/acs.chemrestox.2c00268
    2. Sangram Gore, Kirill Ukhanov, Cyril Herbivo, Naeem Asad, Yuriy V. Bobkov, Jeffrey R. Martens, Timothy M. Dore. Photoactivatable Odorants for Chemosensory Research. ACS Chemical Biology 2020, 15 (9) , 2516-2528. https://doi.org/10.1021/acschembio.0c00541
    3. Igor A. Schepetkin, Svetlana V. Kushnarenko, Gulmira Özek, Liliya N. Kirpotina, Pritam Sinharoy, Gulzhakhan A. Utegenova, Karime T. Abidkulova, Temel Özek, Kemal Hüsnü Can Başer, Anastasia R. Kovrizhina, Andrei I. Khlebnikov, Derek S. Damron, and Mark T. Quinn . Modulation of Human Neutrophil Responses by the Essential Oils from Ferula akitschkensis and Their Constituents. Journal of Agricultural and Food Chemistry 2016, 64 (38) , 7156-7170. https://doi.org/10.1021/acs.jafc.6b03205
    4. Benjamin F. Cravatt . TRP Channels—Convergent Sites of Action for Phytochemicals and Endogenous Lipid Transmitters That Regulate Human Sensation and Physiology. ACS Chemical Neuroscience 2014, 5 (11) , 1083-1083. https://doi.org/10.1021/cn500263c
    5. Songchao Xu, Yun Wang. Transient Receptor Potential Channels: Multiple Modulators of Peripheral Neuropathic Pain in Several Rodent Models. Neurochemical Research 2024, 49 (4) , 872-886. https://doi.org/10.1007/s11064-023-04087-4
    6. Jonathan S. Marchant, . Progress interrogating TRPMPZQ as the target of praziquantel. PLOS Neglected Tropical Diseases 2024, 18 (2) , e0011929. https://doi.org/10.1371/journal.pntd.0011929
    7. Junqing Gao, Huanhuan Li, Hua Lv, Xiansong Cheng. Mutation of TRPML1 Channel and Pathogenesis of Neurodegeneration in Haimeria. Molecular Neurobiology 2023, 12 https://doi.org/10.1007/s12035-023-03874-y
    8. Ebrahim Nasiri, Jamal Rezaei Orimi, Zahra Aghabeiglooei, Kathleen Walker-Meikle, Mohammad Amrollahi-Sharifabadi. Avicenna’s pharmacopeia for the treatment of animal bites. Naunyn-Schmiedeberg's Archives of Pharmacology 2023, 396 (12) , 3375-3393. https://doi.org/10.1007/s00210-023-02586-3
    9. Tzu‐Ho Chen, Hsiao‐Ching Lin. Terpene Synthases in the Biosynthesis of Drimane‐Type Sesquiterpenes across Diverse Organisms. ChemBioChem 2023, 24 (22) https://doi.org/10.1002/cbic.202300518
    10. Solpa Lee, Minwoo Kim, Bum Ju Ahn, Yongwoo Jang. Odorant-responsive biological receptors and electronic noses for volatile organic compounds with aldehyde for human health and diseases: A perspective review. Journal of Hazardous Materials 2023, 455 , 131555. https://doi.org/10.1016/j.jhazmat.2023.131555
    11. Yi-Yue Zhang, Xi-Sheng Li, Kai-Di Ren, Jun Peng, Xiu-Ju Luo. Restoration of metal homeostasis: a potential strategy against neurodegenerative diseases. Ageing Research Reviews 2023, 87 , 101931. https://doi.org/10.1016/j.arr.2023.101931
    12. Jiong Zhang, Min Zhang, Bhesh Bhandari, Mingqi Wang. Basic sensory properties of essential oils from aromatic plants and their applications: a critical review. Critical Reviews in Food Science and Nutrition 2023, 14 , 1-14. https://doi.org/10.1080/10408398.2023.2177611
    13. Eiji Takai, Kohki Nakane, Hiroki Takada. Nonlinear Analyses of Electrogastrogram Measurements Taken During Olfactory Stimulation Altering Autonomic Nerve Activity. 2023, 277-287. https://doi.org/10.1007/978-3-031-35681-0_18
    14. Mohammad Zakir Hossain, Hiroshi Ando, Shumpei Unno, Junichi Kitagawa. TRPA1s act as chemosensors but not as cold sensors or mechanosensors to trigger the swallowing reflex in rats. Scientific Reports 2022, 12 (1) https://doi.org/10.1038/s41598-022-07400-3
    15. Jose F. Cabello-Gómez, María Arántzazu Aguinaga-Casañas, Ana Falcón-Piñeiro, Elías González-Gragera, Raquel Márquez-Martín, María del Mar Agraso, Laura Bermúdez, Alberto Baños, Manuel Martínez-Bueno. Antibacterial and Antiparasitic Activity of Propyl-Propane-Thiosulfinate (PTS) and Propyl-Propane-Thiosulfonate (PTSO) from Allium cepa against Gilthead Sea Bream Pathogens in In Vitro and In Vivo Studies. Molecules 2022, 27 (20) , 6900. https://doi.org/10.3390/molecules27206900
    16. Reena Gupta, Bhupinder Kapoor, Monica Gulati, Bimlesh Kumar, Mukta Gupta, Sachin Kumar Singh, Ankit Awasthi. Sweet pepper and its principle constituent capsiate: functional properties and health benefits. Critical Reviews in Food Science and Nutrition 2022, 62 (26) , 7370-7394. https://doi.org/10.1080/10408398.2021.1913989
    17. Zhongqiang Qi, Lina Li, Cunfa Xu, Muxing Liu, Yousheng Wang, Li Zhang, Jian Chen, Haiyan Lu, Zhiqi Shi. The Sodium/Calcium Exchanger PcNCX1-Mediated Ca2+ Efflux Is Involved in Cinnamaldehyde-Induced Cell-Wall Defects of Phytophthora capsici. Agronomy 2022, 12 (8) , 1763. https://doi.org/10.3390/agronomy12081763
    18. Thomas Mathew, Saji Kaithavalappil John, Mahendra Javali, Manoj Vasireddy, Raghunandan Nadig, Gosala Raja Kukkuta Sarma. Substance use related cluster headache: A case series. Headache: The Journal of Head and Face Pain 2022, 62 (7) , 908-910. https://doi.org/10.1111/head.14364
    19. José Luis Álvarez- Vásquez, Nathaly Fernanda Parra- Solano, Gabriela Elizabeth Saavedra-Cornejo, Ximena Elizabeth Espinosa-Vásquez. Global use of Ethnomedicinal Plants to Treat Toothache. Biomedical and Pharmacology Journal 2022, 15 (2) , 847-881. https://doi.org/10.13005/bpj/2421
    20. Ning Cai, Alvin Chi-Keung Lai, Kin Liao, Peter R. Corridon, David J. Graves, Vincent Chan. Recent Advances in Fluorescence Recovery after Photobleaching for Decoupling Transport and Kinetics of Biomacromolecules in Cellular Physiology. Polymers 2022, 14 (9) , 1913. https://doi.org/10.3390/polym14091913
    21. Mariia Nesterkina, Serhii Smola, Nataliya Rusakova, Iryna Kravchenko. Terpenoid Hydrazones as Biomembrane Penetration Enhancers: FT-IR Spectroscopy and Fluorescence Probe Studies. Molecules 2022, 27 (1) , 206. https://doi.org/10.3390/molecules27010206
    22. Thomas Mathew, Saji Kaithavalappil John, Mahendra Vishwanath Javali. Essential oils and cluster headache: insights from two cases. BMJ Case Reports 2021, 14 (8) , e243812. https://doi.org/10.1136/bcr-2021-243812
    23. Sachiko Koyama, Kenji Kondo, Rumi Ueha, Hideki Kashiwadani, Thomas Heinbockel. Possible Use of Phytochemicals for Recovery from COVID-19-Induced Anosmia and Ageusia. International Journal of Molecular Sciences 2021, 22 (16) , 8912. https://doi.org/10.3390/ijms22168912
    24. David Reyes-Camacho, José F. Pérez, Ester Vinyeta, Tobias Aumiller, Jan D. Van der Klis, David Solà-Oriol. Prenatal Exposure to Innately Preferred D-Limonene and Trans-Anethole Does Not Overcome Innate Aversion to Eucalyptol, Affecting Growth Performance of Weanling Piglets. Animals 2021, 11 (7) , 2062. https://doi.org/10.3390/ani11072062
    25. Junfan Fang, Junying Du, Xuaner Xiang, Xiaomei Shao, Xiaofeng He, Yongliang Jiang, Boyi Liu, Yi Liang, Jianqiao Fang. SNI and CFA induce similar changes in TRPV1 and P2X3 expressions in the acute phase but not in the chronic phase of pain. Experimental Brain Research 2021, 239 (3) , 983-995. https://doi.org/10.1007/s00221-020-05988-4
    26. Joanna K. Bujak, Daria Kosmala, Kinga Majchrzak-Kuligowska, Piotr Bednarczyk. Functional Expression of TRPV1 Ion Channel in the Canine Peripheral Blood Mononuclear Cells. International Journal of Molecular Sciences 2021, 22 (6) , 3177. https://doi.org/10.3390/ijms22063177
    27. Robert Tarran, R Graham Barr, Neal L Benowitz, Aruni Bhatnagar, Hong W Chu, Pamela Dalton, Claire M Doerschuk, M Bradley Drummond, Diane R Gold, Maciej L Goniewicz, Eric R Gross, Nadia N Hansel, Philip K Hopke, Robert A Kloner, Vladimir B Mikheev, Evan W Neczypor, Kent E Pinkerton, Lisa Postow, Irfan Rahman, Jonathan M Samet, Matthias Salathe, Catherine M Stoney, Philip S Tsao, Rachel Widome, Tian Xia, DaLiao Xiao, Loren E Wold. E-Cigarettes and Cardiopulmonary Health. Function 2021, 2 (2) https://doi.org/10.1093/function/zqab004
    28. Xinlu Ren, Runan Yang, Lin Li, Xiumei Xu, Shangdong Liang. Long non coding RNAs involved in MAPK pathway mechanism mediates diabetic neuropathic pain. Cell Biology International 2020, 44 (12) , 2372-2379. https://doi.org/10.1002/cbin.11457
    29. Akira Murakami. Hormesis-Mediated Mechanisms Underlying Bioactivities of Phytochemicals. Current Pharmacology Reports 2020, 6 (6) , 325-334. https://doi.org/10.1007/s40495-020-00235-4
    30. Monica Ghosh, Igor A. Schepetkin, Gulmira Özek, Temel Özek, Andrei I. Khlebnikov, Derek S. Damron, Mark T. Quinn. Essential Oils from Monarda fistulosa: Chemical Composition and Activation of Transient Receptor Potential A1 (TRPA1) Channels. Molecules 2020, 25 (21) , 4873. https://doi.org/10.3390/molecules25214873
    31. Sarah Mazzotta, Gabriele Carullo, Aniello Schiano Moriello, Pietro Amodeo, Vincenzo Di Marzo, Margarita Vega-Holm, Rosa Maria Vitale, Francesca Aiello, Antonella Brizzi, Luciano De Petrocellis. Design, Synthesis and In Vitro Experimental Validation of Novel TRPV4 Antagonists Inspired by Labdane Diterpenes. Marine Drugs 2020, 18 (10) , 519. https://doi.org/10.3390/md18100519
    32. Robin Herbrechter, Leopoldo R. Beltrán, Paul M. Ziemba, Sascha Titt, Konstantin Lashuk, André Gottemeyer, Janina Levermann, Katrin M. Hoffmann, Madeline Beltrán, Hanns Hatt, Klemens F. Störtkuhl, Markus Werner, Günter Gisselmann. Effect of 158 herbal remedies on human TRPV1 and the two-pore domain potassium channels KCNK2, 3 and 9. Journal of Traditional and Complementary Medicine 2020, 10 (5) , 446-453. https://doi.org/10.1016/j.jtcme.2020.04.005
    33. Mohammad Zakir Hossain, Hiroshi Ando, Shumpei Unno, Junichi Kitagawa. Targeting Chemosensory Ion Channels in Peripheral Swallowing-Related Regions for the Management of Oropharyngeal Dysphagia. International Journal of Molecular Sciences 2020, 21 (17) , 6214. https://doi.org/10.3390/ijms21176214
    34. Andrea Toschi, Barbara Rossi, Benedetta Tugnoli, Andrea Piva, Ester Grilli. Nature-Identical Compounds and Organic Acids Ameliorate and Prevent the Damages Induced by an Inflammatory Challenge in Caco-2 Cell Culture. Molecules 2020, 25 (18) , 4296. https://doi.org/10.3390/molecules25184296
    35. Eszter Csikós, Kata Csekő, Amir Reza Ashraf, Ágnes Kemény, László Kereskai, Béla Kocsis, Andrea Böszörményi, Zsuzsanna Helyes, Györgyi Horváth. Effects of Thymus vulgaris L., Cinnamomum verum J.Presl and Cymbopogon nardus (L.) Rendle Essential Oils in the Endotoxin-induced Acute Airway Inflammation Mouse Model. Molecules 2020, 25 (15) , 3553. https://doi.org/10.3390/molecules25153553
    36. Mariia Nesterkina, Luidmyla Ognichenko, Angela Shyrykalova, Iryna Kravchenko, Victor Kuz’min. QSAR models for analgesic activity prediction of terpenes and their derivatives. Structural Chemistry 2020, 31 (3) , 947-954. https://doi.org/10.1007/s11224-019-01479-7
    37. Antonella Brizzi, Francesca Aiello, Serena Boccella, Maria Grazia Cascio, Luciano De Petrocellis, Maria Frosini, Francesca Gado, Alessia Ligresti, Livio Luongo, Pietro Marini, Claudia Mugnaini, Federica Pessina, Federico Corelli, Sabatino Maione, Clementina Manera, Roger G. Pertwee, Vincenzo Di Marzo. Synthetic bioactive olivetol-related amides: The influence of the phenolic group in cannabinoid receptor activity. Bioorganic & Medicinal Chemistry 2020, 28 (11) , 115513. https://doi.org/10.1016/j.bmc.2020.115513
    38. Chris McKennan, Carole Ober, Dan Nicolae. Estimation and inference in metabolomics with nonrandom missing data and latent factors. The Annals of Applied Statistics 2020, 14 (2) https://doi.org/10.1214/20-AOAS1328
    39. Filomena Perri, Adriana Coricello, James D. Adams. Monoterpenoids: The Next Frontier in the Treatment of Chronic Pain?. J — Multidisciplinary Scientific Journal 2020, 3 (2) , 195-214. https://doi.org/10.3390/j3020016
    40. Allisson Benatti Justino, Marilia Fontes Barbosa, Thiago Vieira Neves, Heitor Cappato Guerra Silva, Evelyne da Silva Brum, Maria Fernanda Pessano Fialho, Ana Cláudia Couto, André Lopes Saraiva, Veridiana de Melo Rodrigues Avila, Sara Marchesan Oliveira, Marcos Pivatto, Foued Salmen Espindola, Cassia Regina Silva. Stephalagine, an aporphine alkaloid from Annona crassiflora fruit peel, induces antinociceptive effects by TRPA1 and TRPV1 channels modulation in mice. Bioorganic Chemistry 2020, 96 , 103562. https://doi.org/10.1016/j.bioorg.2019.103562
    41. Sachiko Koyama, Thomas Heinbockel. The Effects of Essential Oils and Terpenes in Relation to Their Routes of Intake and Application. International Journal of Molecular Sciences 2020, 21 (5) , 1558. https://doi.org/10.3390/ijms21051558
    42. Sachiko Koyama, Anna Purk, Manpreet Kaur, Helena A. Soini, Milos V. Novotny, Keith Davis, C. Cheng Kao, Hiroaki Matsunami, Anthony Mescher, . Beta-caryophyllene enhances wound healing through multiple routes. PLOS ONE 2019, 14 (12) , e0216104. https://doi.org/10.1371/journal.pone.0216104
    43. Yu. A. Boiko, M. V. Nesterkina, A. A. Shandra, I. A. Kravchenko. Analgesic and Anti-Inflammatory Activity of Vanillin Derivatives. Pharmaceutical Chemistry Journal 2019, 53 (7) , 650-654. https://doi.org/10.1007/s11094-019-02056-2
    44. Anne E. Turco, Mark T. Cadena, Helen L. Zhang, Jaskiran K. Sandhu, Steven R. Oakes, Thrishna Chathurvedula, Richard E. Peterson, Janet R. Keast, Chad M. Vezina. A temporal and spatial map of axons in developing mouse prostate. Histochemistry and Cell Biology 2019, 152 (1) , 35-45. https://doi.org/10.1007/s00418-019-01784-6
    45. Bruna Benso, Daniel Bustos, Miguel O. Zarraga, Wendy Gonzalez, Julio Caballero, Sebastian Brauchi. Chalcone derivatives as non-canonical ligands of TRPV1. The International Journal of Biochemistry & Cell Biology 2019, 112 , 18-23. https://doi.org/10.1016/j.biocel.2019.04.010
    46. Jisun Oh, Chan Ho Jang, Jong-Sang Kim. Soy-derived phytoalexins: mechanism of in vivo biological effectiveness in spite of their low bioavailability. Food Science and Biotechnology 2019, 28 (1) , 1-6. https://doi.org/10.1007/s10068-018-0498-7
    47. C. Jansen, L.M.N Shimoda, J.K. Kawakami, L. Ang, A.J. Bacani, J.D. Baker, C. Badowski, M. Speck, A.J. Stokes, A.L. Small-Howard, H Turner. Myrcene and terpene regulation of TRPV1. Channels 2019, 13 (1) , 344-366. https://doi.org/10.1080/19336950.2019.1654347
    48. Pengfei Han, Stephanie Mann, Claudia Raue, Jonathan Warr, Thomas Hummel. Pepper with and without a sting: Brain processing of intranasal trigeminal and olfactory stimuli from the same source. Brain Research 2018, 1700 , 41-46. https://doi.org/10.1016/j.brainres.2018.07.010
    49. M. Bishnoi, P. Khare, L. Brown, S. K. Panchal. Transient receptor potential (TRP) channels: a metabolic TR(i)P to obesity prevention and therapy. Obesity Reviews 2018, 19 (9) , 1269-1292. https://doi.org/10.1111/obr.12703
    50. M. Flori Sassano, Eric S. Davis, James E. Keating, Bryan T. Zorn, Tavleen K. Kochar, Matthew C. Wolfgang, Gary L. Glish, Robert Tarran, . Evaluation of e-liquid toxicity using an open-source high-throughput screening assay. PLOS Biology 2018, 16 (3) , e2003904. https://doi.org/10.1371/journal.pbio.2003904
    51. Chenglong Liu, Congcong Li, Zeyu Deng, Errong Du, Changshui Xu. Long Non-coding RNA BC168687 is Involved in TRPV1-mediated Diabetic Neuropathic Pain in Rats. Neuroscience 2018, 374 , 214-222. https://doi.org/10.1016/j.neuroscience.2018.01.049
    52. Phillipa K. Beale, Karen J. Marsh, William J. Foley, Ben D. Moore. A hot lunch for herbivores: physiological effects of elevated temperatures on mammalian feeding ecology. Biological Reviews 2018, 93 (1) , 674-692. https://doi.org/10.1111/brv.12364
    53. Angelika Böttger, Ute Vothknecht, Cordelia Bolle, Alexander Wolf. Plant-Derived Drugs Affecting Ion Channels. 2018, 121-140. https://doi.org/10.1007/978-3-319-99546-5_8
    54. Mariia Nesterkina, Iryna Kravchenko. Synthesis and Pharmacological Properties of Novel Esters Based on Monoterpenoids and Glycine. Pharmaceuticals 2017, 10 (4) , 47. https://doi.org/10.3390/ph10020047
    55. Phillip W. Clapp, Ilona Jaspers. Electronic Cigarettes: Their Constituents and Potential Links to Asthma. Current Allergy and Asthma Reports 2017, 17 (11) https://doi.org/10.1007/s11882-017-0747-5
    56. Gulmira Oüzek, Igor A Schepetkin, Gulzhakhan A Utegenova, Liliya N Kirpotina, Spencer R Andrei, Temel Oüzek, Kemal Huüsnuü Can Baser, Karime T Abidkulova, Svetlana V Kushnarenko, Andrei I Khlebnikov, Derek S Damron, Mark T Quinn. Chemical composition and phagocyte immunomodulatory activity of Ferula iliensis essential oils. Journal of Leukocyte Biology 2017, 101 (6) , 1361-1371. https://doi.org/10.1189/jlb.3A1216-518RR
    57. Kristina Friedland, Christian Harteneck. Spices and Odorants as TRP Channel Activators. 2017, 85-86. https://doi.org/10.1007/978-3-319-26932-0_34
    58. Mariia Nesterkina, Iryna Kravchenko. Synthesis and Pharmacological Properties of Novel Esters Based on Monocyclic Terpenes and GABA. Pharmaceuticals 2016, 9 (2) , 32. https://doi.org/10.3390/ph9020032
    59. James Adams. The Effects of Yin, Yang and Qi in the Skin on Pain. Medicines 2016, 3 (1) , 5. https://doi.org/10.3390/medicines3010005
    60. Brett Boonen, Justyna B. Startek, Karel Talavera. Chemical Activation of Sensory TRP Channels. 2016, 73-113. https://doi.org/10.1007/7355_2015_98
    • Abstract

      Figure 1

      Figure 1. TRPA1 agonists (phytochemicals, synthetic chemicals, and endogenous molecules). Allicin, 2-propene-1-sulfinothioic acid S-2-propenyl ester; allylisothiocyanate (AITC), 3-isothiocyanato-1-propene, is an organosulfur compound; cinnamaldehyde, (2E)-3-phenylprop-2-enal; Δ9-tetrahydrocannabinol, (−)-(6aR,10aR)-6,6,9-trimethyl-3-pentyl-6a,7,8,10a-tetrahydro-6H-benzo[c]chromen-1-ol; umbellulone, 1-isopropyl-4-methylbicyclo[3.1.0]hex-3-en-2-one; ligustilide, (3Z)-3-butylidene-4,5-dihydro-2-benzofuran-1(3H)-one; methyl salicylate, methyl 2-hydroxybenzoate; icilin, 1-(2-hydroxyphenyl)-4-(3-nitrophenyl)-3,6-dihydropyrimidin-2-one; N-methylmaleimide (oxidizing agent); 4-hydroxynonenal, 4-hydroxy-2-nonenal, an α,β-unsaturated hydroxyalkenal produced by lipid peroxidation, is an endogenous agonist; methylglyoxal, an aldehyde from pyruvic acid, acts both as an aldehyde and ketone, and reacts with free amino acids such as lysine, arginine and thiol groups of cysteine. MG is an endogenous agonist.

      Figure 2

      Figure 2. Other TRP channel activators. Hyperforin, (1R,5S,6R,7S)-4-hydroxy-5-isobutyryl-6-methyl-1,3,7-tris(3-methyl-2-buten-1-yl)-6-(4-methyl-3-penten-1-yl)bicyclo[3.3.1]non-3-ene-2,9-dione (TRPC6); bisandrographolide, 3-{(E)-2-[6-hydroxy-5-(hydroxymethyl)-5,8a-dimethyl-2-methylenedecahydro-1-naphthalenyl]vinyl}-5-{6-hydroxy-5-(hydroxymethyl)-5,8a-dimethyl-2-methylene-1-[2-(2-oxo-2,5-dihydro-3-furanyl)ethyl]decahydro-1-naphthalenyl}-2(5H)-furanone (TRPV4); paclitaxel, (2α,5β,7β,10β,13α)-4,10-diacetoxy-13-{[(2R,3S)-3-(benzoylamino)-2-hydroxy-3-phenylpropanoyl]oxy}-1,7-dihydroxy-9-oxo-5,20-epoxytax-11-en-2-yl benzoate (TRPA1, TRPV4); dicentrine, (7aS)-10,11-dimethoxy-7-methyl-6,7,7a,8-tetrahydro-5H-[1,3]benzodioxolo[6,5,4-de]benzo[g]quinoline (TRPA1).

      Figure 3

      Figure 3. TRPM8 agonists. Menthol, (1R,2S,5R)-2-isopropyl-5-methylcyclohexanol; eucalyptol, 1,3,3-trimethyl-2-oxabicyclo[2.2.2]octane; icilin, 1-(2-hydroxyphenyl)-4-(3-nitrophenyl)-3,6-dihydropyrimidin-2-one.

      Figure 4

      Figure 4. TRPV1 agonists. Vanillin, has the vanillyl moiety, that is essential for activating TRPV1 channels; Capsaicin, (E)-N-[(4-hydroxy-3-methoxyphenyl)methyl]-8-methyl-6-nonenamide, has a vanilloid and an aceyl moiety; dihydrocapsaicin, the structure of which is a 6,7-dihydro derivative of capsaicin; resiniferatoxin, has a complex structure, but shares a homovanillyl group, which is necessary for the activity of all vanilloids; eugenol, 2-methoxy-4-(2-propenyl)phenol and is a member of the allylbenzene class of chemical compounds; Cannabidiol, 2-[(1R,6R)-6-isopropenyl-3-methyl-3-cyclohexen-1-yl]-5-pentyl-1,3-benzenediol; anandamide or arachidonylethanolamide or arachidonic acid N-(hydroxyethyl)amide consists of the acyl moiety and is an edogenous ligand of TRPV1 and cannabinoid receptor 1 (CB1).

    • References

      ARTICLE SECTIONS
      Jump To

      This article references 159 other publications.

      1. 1
        Caterina, M. J., Schumacher, M. A., Tominaga, M., Rosen, T. A., Levine, J. D., and Julius, D. (1997) The capsaicin receptor: a heat-activated ion channel in the pain pathway Nature 389, 816 824
      2. 2
        Cosens, D. J. and Manning, A. (1969) Abnormal electroretinogram from a Drosophila mutant Nature 224, 285 287
      3. 3
        Montell, C. and Rubin, G. M. (1989) Molecular characterization of the Drosophila trp locus: a putative integral membrane protein required for phototransduction Neuron 2, 1313 1323
      4. 4
        Montell, C. (2005) The TRP superfamily of cation channels Sci. STKE re3
      5. 5
        Story, G. M., Peier, A. M., Reeve, A. J., Eid, S. R., Mosbacher, J., Hricik, T. R., Earley, T. J., Hergarden, A. C., Andersson, D. A., Hwang, S. W., McIntyre, P., Jegla, T., Bevan, S., and Patapoutian, A. (2003) ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures Cell 112, 819 829
      6. 6
        Jordt, S. E., Bautista, D. M., Chuang, H. H., McKemy, D. D., Zygmunt, P. M., Hogestatt, E. D., Meng, I. D., and Julius, D. (2004) Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1 Nature 427, 260 265
      7. 7
        Peier, A. M., Moqrich, A., Hergarden, A. C., Reeve, A. J., Andersson, D. A., Story, G. M., Earley, T. J., Dragoni, I., McIntyre, P., Bevan, S., and Patapoutian, A. (2002) A TRP channel that senses cold stimuli and menthol Cell 108, 705 715
      8. 8
        McKemy, D. D., Neuhausser, W. M., and Julius, D. (2002) Identification of a cold receptor reveals a general role for TRP channels in thermosensation Nature 416, 52 58
      9. 9
        Cao, E., Liao, M., Cheng, Y., and Julius, D. (2013) TRPV1 structures in distinct conformations reveal activation mechanisms Nature 504, 113 118
      10. 10
        Liao, M., Cao, E., Julius, D., and Cheng, Y. (2013) Structure of the TRPV1 ion channel determined by electron cryo-microscopy Nature 504, 107 112
      11. 11
        Bikman, B. T., Zheng, D., Pories, W. J., Chapman, W., Pender, J. R., Bowden, R. C., Reed, M. A., Cortright, R. N., Tapscott, E. B., Houmard, J. A., Tanner, C. J., Lee, J., and Dohm, G. L. (2008) Mechanism for improved insulin sensitivity after gastric bypass surgery J. Clin. Endocrinol. Metab. 93, 4656 4663
      12. 12
        Corey, D. P., Garcia-Anoveros, J., Holt, J. R., Kwan, K. Y., Lin, S. Y., Vollrath, M. A., Amalfitano, A., Cheung, E. L., Derfler, B. H., Duggan, A., Geleoc, G. S., Gray, P. A., Hoffman, M. P., Rehm, H. L., Tamasauskas, D., and Zhang, D. S. (2004) TRPA1 is a candidate for the mechanosensitive transduction channel of vertebrate hair cells Nature 432, 723 730
      13. 13
        Bandell, M., Story, G. M., Hwang, S. W., Viswanath, V., Eid, S. R., Petrus, M. J., Earley, T. J., Patapoutian, A., Bautista, D. M., Movahed, P., Hinman, A., Axelsson, H. E., Sterner, O., Hogestatt, E. D., Julius, D., Jordt, S. E., and Zygmunt, P. M. (2004) Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin Neuron 41, 849 857
      14. 14
        Bautista, D. M., Movahed, P., Hinman, A., Axelsson, H. E., Sterner, O., Hogestatt, E. D., Julius, D., Jordt, S. E., and Zygmunt, P. M. (2005) Pungent products from garlic activate the sensory ion channel TRPA1 Proc. Natl. Acad. Sci. U.S.A. 102, 12248 12252
      15. 15
        Macpherson, L. J., Geierstanger, B. H., Viswanath, V., Bandell, M., Eid, S. R., Hwang, S., and Patapoutian, A. (2005) The pungency of garlic: activation of TRPA1 and TRPV1 in response to allicin Curr. Biol. 15, 929 934
      16. 16
        Kwan, K. Y., Allchorne, A. J., Vollrath, M. A., Christensen, A. P., Zhang, D. S., Woolf, C. J., and Corey, D. P. (2006) TRPA1 contributes to cold, mechanical, and chemical nociception but is not essential for hair-cell transduction Neuron 50, 277 289
      17. 17
        Andersson, D. A., Gentry, C., Moss, S., and Bevan, S. (2008) Transient receptor potential A1 is a sensory receptor for multiple products of oxidative stress J. Neurosci. 28, 2485 2494
      18. 18
        Bessac, B. F., Sivula, M., von Hehn, C. A., Escalera, J., Cohn, L., and Jordt, S. E. (2008) TRPA1 is a major oxidant sensor in murine airway sensory neurons J. Clin. Invest. 118, 1899 1910
      19. 19
        Cao, D. S., Zhong, L., Hsieh, T. H., Abooj, M., Bishnoi, M., Hughes, L., and Premkumar, L. S. (2012) Expression of transient receptor potential ankyrin 1 (TRPA1) and its role in insulin release from rat pancreatic beta cells PloS One 7, e38005
      20. 20
        Brownlee, M. (2001) Biochemistry and molecular cell biology of diabetic complications Nature 414, 813 820
      21. 21
        Nagata, K., Duggan, A., Kumar, G., and Garcia-Anoveros, J. (2005) Nociceptor and hair cell transducer properties of TRPA1, a channel for pain and hearing J. Neurosci. 25, 4052 4061
      22. 22
        Obata, K., Katsura, H., Mizushima, T., Yamanaka, H., Kobayashi, K., Dai, Y., Fukuoka, T., Tokunaga, A., Tominaga, M., and Noguchi, K. (2005) TRPA1 induced in sensory neurons contributes to cold hyperalgesia after inflammation and nerve injury J. Clin. Invest. 115, 2393 2401
      23. 23
        Hinman, A., Chuang, H. H., Bautista, D. M., and Julius, D. (2006) TRP channel activation by reversible covalent modification Proc. Natl. Acad. Sci. U.S.A. 103, 19564 19568
      24. 24
        Raisinghani, M., Zhong, L., Jeffry, J. A., Bishnoi, M., Pabbidi, R. M., Pimentel, F., Cao, D. S., Evans, M. S., and Premkumar, L. S. (2011) Activation characteristics of transient receptor potential ankyrin 1 and its role in nociception Am. J. Physiol.: Cell Physiol. 301, C587 600
      25. 25
        Premkumar, L. S. and Raisinghani, M. (2006) Nociceptors in cardiovascular functions: complex interplay as a result of cyclooxygenase inhibition Mol. Pain 2, 26
      26. 26
        Khan, A., Safdar, M., Ali Khan, M. M., Khattak, K. N., and Anderson, R. A. (2003) Cinnamon improves glucose and lipids of people with type 2 diabetes Diabetes Care 26, 3215 3218
      27. 27
        Hlebowicz, J., Darwiche, G., Bjorgell, O., and Almer, L. O. (2007) Effect of cinnamon on postprandial blood glucose, gastric emptying, and satiety in healthy subjects Am. J. Clin. Nutr. 85, 1552 1556
      28. 28
        Lissiman, E., Bhasale, A. L., and Cohen, M. (2009) Garlic for the common cold Cochrane Database Syst. Rev. 3, CD006206
      29. 29
        Lissiman, E., Bhasale, A. L., and Cohen, M. (2012) Garlic for the common cold Cochrane Database Syst. Rev. 14, CD006206
      30. 30
        Cutler, R. R., Odent, M., Hajj-Ahmad, H., Maharjan, S., Bennett, N. J., Josling, P. D., Ball, V., Hatton, P., and Dall’Antonia, M. (2009) In vitro activity of an aqueous allicin extract and a novel allicin topical gel formulation against Lancefield group B streptococci J. Antimicrob. Chemother. 63, 151 154
      31. 31
        Khanum, F., Anilakumar, K. R., and Viswanathan, K. R. (2004) Anticarcinogenic properties of garlic: a review Crit. Rev. Food. Sci. Nutr. 44, 479 488
      32. 32
        Shenoy, N. R. and Choughuley, A. S. (1992) Inhibitory effect of diet related sulphydryl compounds on the formation of carcinogenic nitrosamines Cancer Lett. 65, 227 232
      33. 33
        Koizumi, K., Iwasaki, Y., Narukawa, M., Iitsuka, Y., Fukao, T., Seki, T., Ariga, T., and Watanabe, T. (2009) Diallyl sulfides in garlic activate both TRPA1 and TRPV1 Biochem. Biophys. Res. Commun. 382, 545 548
      34. 34
        Leamy, A. W., Shukla, P., McAlexander, M. A., Carr, M. J., and Ghatta, S. (2011) Curcumin ((E,E)-1,7-bis(4-hydroxy-3-methoxyphenyl)-1,6-heptadiene-3,5-dione) activates and desensitizes the nociceptor ion channel TRPA1 Neurosci. Lett. 503, 157 162
      35. 35
        Gupta, S. C., Patchva, S., and Aggarwal, B. B. (2013) Therapeutic roles of curcumin: lessons learned from clinical trials AAPS J. 15, 195 218
      36. 36
        Song, Y., Sonawane, N. D., Salinas, D., Qian, L., Pedemonte, N., Galietta, L. J., and Verkman, A. S. (2004) Evidence against the rescue of defective DeltaF508-CFTR cellular processing by curcumin in cell culture and mouse models J. Biol. Chem. 279, 40629 40633
      37. 37
        Gao, J., Zhou, H., Lei, T., Zhou, L., Li, W., Li, X., and Yang, B. (2011) Curcumin inhibits renal cyst formation and enlargement in vitro by regulating intracellular signaling pathways Eur. J. Pharmacol. 654, 92 99
      38. 38
        Sohma, Y., Yu, Y. C., and Hwang, T. C. (2013) Curcumin and genistein: the combined effects on disease-associated CFTR mutants and their clinical implications Curr. Pharm. Des. 19, 3521 3528
      39. 39
        Nassini, R., Materazzi, S., Vriens, J., Prenen, J., Benemei, S., De Siena, G., la Marca, G., Andre, E., Preti, D., Avonto, C., Sadofsky, L., Di Marzo, V., De Petrocellis, L., Dussor, G., Porreca, F., Taglialatela-Scafati, O., Appendino, G., Nilius, B., and Geppetti, P. (2012) The ‘headache tree’ via umbellulone and TRPA1 activates the trigeminovascular system Brain 135, 376 390
      40. 40
        Zhong, J., Pollastro, F., Prenen, J., Zhu, Z., Appendino, G., and Nilius, B. (2011) Ligustilide: a novel TRPA1 modulator Pfluegers Arch. 462, 841 849
      41. 41
        Materazzi, S., Fusi, C., Benemei, S., Pedretti, P., Patacchini, R., Nilius, B., Prenen, J., Creminon, C., Geppetti, P., and Nassini, R. (2012) TRPA1 and TRPV4 mediate paclitaxel-induced peripheral neuropathy in mice via a glutathione-sensitive mechanism Pfluegers Arch. 463, 561 569
      42. 42
        De Petrocellis, L., Vellani, V., Schiano-Moriello, A., Marini, P., Magherini, P. C., Orlando, P., and Di Marzo, V. (2008) Plant-derived cannabinoids modulate the activity of transient receptor potential channels of ankyrin type-1 and melastatin type-8 J. Pharmacol. Exp. Ther. 325, 1007 1015
      43. 43
        Montrucchio, D. P., Cordova, M. M., and Santos, A. R. (2013) Plant derived aporphinic alkaloid S-(+)-dicentrine induces antinociceptive effect in both acute and chronic inflammatory pain models: evidence for a role of TRPA1 channels PloS One 8, e67730
      44. 44
        Talavera, K., Gees, M., Karashima, Y., Meseguer, V. M., Vanoirbeek, J. A., Damann, N., Everaerts, W., Benoit, M., Janssens, A., Vennekens, R., Viana, F., Nemery, B., Nilius, B., and Voets, T. (2009) Nicotine activates the chemosensory cation channel TRPA1 Nat. Neurosci. 12, 1293 1299
      45. 45
        Riera, C. E., Menozzi-Smarrito, C., Affolter, M., Michlig, S., Munari, C., Robert, F., Vogel, H., Simon, S. A., and Le Coutre, J. (2009) Compounds from Sichuan and Melegueta peppers activate, covalently and non-covalently, TRPA1 and TRPV1 channels Br. J. Pharmacol. 157, 1398 1409
      46. 46
        Vazquez, G., Wedel, B. J., Aziz, O., Trebak, M., and Putney, J. W., Jr. (2004) The mammalian TRPC cation channels Biochim. Biophys. Acta 1742, 21 36
      47. 47
        Birnbaumer, L. (2009) The TRPC class of ion channels: a critical review of their roles in slow, sustained increases in intracellular Ca(2+) concentrations Annu. Rev. Pharmacol. Toxicol. 49, 395 426
      48. 48
        Leuner, K., Kazanski, V., Muller, M., Essin, K., Henke, B., Gollasch, M., Harteneck, C., and Muller, W. E. (2007) Hyperforin--a key constituent of St. John’s wort specifically activates TRPC6 channels FASEB J. 21, 4101 4111
      49. 49
        Sikand, P. and Premkumar, L. S. (2007) Potentiation of glutamatergic synaptic transmission by protein kinase C-mediated sensitization of TRPV1 at the first sensory synapse J. Physiol. 581, 631 647
      50. 50
        Cao, D. S., Yu, S. Q., and Premkumar, L. S. (2009) Modulation of transient receptor potential Vanilloid 4-mediated membrane currents and synaptic transmission by protein kinase C Mol. Pain 5, 5
      51. 51
        Jeffry, J. A., Yu, S. Q., Sikand, P., Parihar, A., Evans, M. S., and Premkumar, L. S. (2009) Selective targeting of TRPV1 expressing sensory nerve terminals in the spinal cord for long lasting analgesia PloS One 4, e7021
      52. 52
        Evans, M. S., Cheng, X., Jeffry, J. A., Disney, K. E., and Premkumar, L. S. (2012) Sumatriptan inhibits TRPV1 channels in trigeminal neurons Headache 52, 773 784
      53. 53
        Zhou, J., Du, W., Zhou, K., Tai, Y., Yao, H., Jia, Y., and Ding, Y. (2008) Critical role of TRPC6 channels in the formation of excitatory synapses Nat. Neurosci. 11, 741 743
      54. 54
        Nilius, B. and Owsianik, G. (2011) The transient receptor potential family of ion channels Genome Biol. 12, 218
      55. 55
        Carakostas, M. C., Curry, L. L., Boileau, A. C., and Brusick, D. J. (2008) Overview: the history, technical function and safety of rebaudioside A, a naturally occurring steviol glycoside, for use in food and beverages Food Chem. Toxicol. 46 (Suppl 7) S1 S10
      56. 56
        Medler, K. F. (2011) Multiple roles for TRPs in the taste system: not your typical TRPs Adv. Exp. Med. Biol. 704, 831 846
      57. 57
        Sprous, D. and Palmer, K. R. (2010) The T1R2/T1R3 sweet receptor and TRPM5 ion channel taste targets with therapeutic potential Prog. Mol. Biol. Transl. Sci. 91, 151 208
      58. 58
        Palmer, R. K. (2007) The pharmacology and signaling of bitter, sweet, and umami taste sensing Mol. Interv. 7, 87 98
      59. 59
        Depoortere, I. (2014) Taste receptors of the gut: emerging roles in health and disease Gut 63, 179 190
      60. 60
        Kaji, I., Karaki, S. I., and Kuwahara, A. (2014) Taste Sensing in the Colon Curr. Pharm. Des. 20, 2766 2774
      61. 61
        Premkumar, L. S., Raisinghani, M., Pingle, S. C., Long, C., and Pimentel, F. (2005) Downregulation of transient receptor potential melastatin 8 by protein kinase C-mediated dephosphorylation J. Neurosci. 25, 11322 11329
      62. 62
        Tsuzuki, K., Xing, H., Ling, J., and Gu, J. G. (2004) Menthol-induced Ca2+ release from presynaptic Ca2+ stores potentiates sensory synaptic transmission J. Neurosci. 24, 762 771
      63. 63
        Tsavaler, L., Shapero, M. H., Morkowski, S., and Laus, R. (2001) Trp-p8, a novel prostate-specific gene, is up-regulated in prostate cancer and other malignancies and shares high homology with transient receptor potential calcium channel proteins Cancer Res. 61, 3760 3769
      64. 64
        Uhl, G. R., Walther, D., Behm, F. M., and Rose, J. E. (2011) Menthol preference among smokers: association with TRPA1 variants Nicotine Tob. Res. 13, 1311 1315
      65. 65
        Hans, M., Wilhelm, M., and Swandulla, D. (2012) Menthol suppresses nicotinic acetylcholine receptor functioning in sensory neurons via allosteric modulation Chem. Senses 37, 463 469
      66. 66
        Willis, D. N., Liu, B., Ha, M. A., Jordt, S. E., and Morris, J. B. (2011) Menthol attenuates respiratory irritation responses to multiple cigarette smoke irritants FASEB J. 25, 4434 4444
      67. 67
        Leuenroth, S. J., Okuhara, D., Shotwell, J. D., Markowitz, G. S., Yu, Z., Somlo, S., and Crews, C. M. (2007) Triptolide is a traditional Chinese medicine-derived inhibitor of polycystic kidney disease Proc. Natl. Acad. Sci. U.S.A. 104, 4389 4394
      68. 68
        Szallasi, A. and Blumberg, P. M. (1999) Vanilloid (Capsaicin) receptors and mechanisms Pharmacol. Rev. 51, 159 212
      69. 69
        Vennekens, R., Owsianik, G., and Nilius, B. (2008) Vanilloid transient receptor potential cation channels: an overview Curr. Pharm. Des. 14, 18 31
      70. 70
        Vriens, J., Nilius, B., and Vennekens, R. (2008) Herbal compounds and toxins modulating TRP channels Curr. Neuropharmacol. 6, 79 96
      71. 71
        Hellwig, N., Albrecht, N., Harteneck, C., Schultz, G., and Schaefer, M. (2005) Homo- and heteromeric assembly of TRPV channel subunits J. Cell Sci. 118, 917 928
      72. 72
        Tominaga, M., Caterina, M. J., Malmberg, A. B., Rosen, T. A., Gilbert, H., Skinner, K., Raumann, B. E., Basbaum, A. I., and Julius, D. (1998) The cloned capsaicin receptor integrates multiple pain-producing stimuli Neuron 21, 531 543
      73. 73
        Nakatsuka, T., Furue, H., Yoshimura, M., and Gu, J. G. (2002) Activation of central terminal vanilloid receptor-1 receptors and alpha beta-methylene-ATP-sensitive P2X receptors reveals a converged synaptic activity onto the deep dorsal horn neurons of the spinal cord J. Neurosci. 22, 1228 1237
      74. 74
        Baccei, M. L., Bardoni, R., and Fitzgerald, M. (2003) Development of nociceptive synaptic inputs to the neonatal rat dorsal horn: glutamate release by capsaicin and menthol J. Physiol. 549, 231 242
      75. 75
        Kim, Y. H., Back, S. K., Davies, A. J., Jeong, H., Jo, H. J., Chung, G., Na, H. S., Bae, Y. C., Kim, S. J., Kim, J. S., Jung, S. J., and Oh, S. B. (2012) TRPV1 in GABAergic interneurons mediates neuropathic mechanical allodynia and disinhibition of the nociceptive circuitry in the spinal cord Neuron 74, 640 647
      76. 76
        Doyle, M. W., Bailey, T. W., Jin, Y. H., and Andresen, M. C. (2002) Vanilloid receptors presynaptically modulate cranial visceral afferent synaptic transmission in nucleus tractus solitarius J. Neurosci. 22, 8222 8229
      77. 77
        Marinelli, S., Di Marzo, V., Berretta, N., Matias, I., Maccarrone, M., Bernardi, G., and Mercuri, N. B. (2003) Presynaptic facilitation of glutamatergic synapses to dopaminergic neurons of the rat substantia nigra by endogenous stimulation of vanilloid receptors J. Neurosci. 23, 3136 3144
      78. 78
        Marinelli, S., Vaughan, C. W., Christie, M. J., and Connor, M. (2002) Capsaicin activation of glutamatergic synaptic transmission in the rat locus coeruleus in vitro J. Physiol. 543, 531 540
      79. 79
        Gibson, H. E., Edwards, J. G., Page, R. S., Van Hook, M. J., and Kauer, J. A. (2008) TRPV1 channels mediate long-term depression at synapses on hippocampal interneurons Neuron 57, 746 759
      80. 80
        Grueter, B. A., Brasnjo, G., and Malenka, R. C. (2010) Postsynaptic TRPV1 triggers cell type-specific long-term depression in the nucleus accumbens Nat. Neurosci. 13, 1519 1525
      81. 81
        Premkumar, L. S. and Ahern, G. P. (2000) Induction of vanilloid receptor channel activity by protein kinase C Nature 408, 985 990
      82. 82
        Caterina, M. J. and Julius, D. (2001) The vanilloid receptor: a molecular gateway to the pain pathway Annu. Rev. Neurosci. 24, 487 517
      83. 83
        Minke, B. and Cook, B. (2002) TRP channel proteins and signal transduction Physiol. Rev. 82, 429 472
      84. 84
        Clapham, D. E. (2003) TRP channels as cellular sensors Nature 426, 517 524
      85. 85
        Premkumar, L. S. and Bishnoi, M. (2011) Disease-related changes in TRPV1 expression and its implications for drug development Curr. Top. Med. Chem. 11, 2192 2209
      86. 86
        Julius, D. (2013) TRP channels and pain Annu. Rev. Cell Dev. Biol. 29, 355 384
      87. 87
        Mezey, E., Toth, Z. E., Cortright, D. N., Arzubi, M. K., Krause, J. E., Elde, R., Guo, A., Blumberg, P. M., and Szallasi, A. (2000) Distribution of mRNA for vanilloid receptor subtype 1 (VR1), and VR1-like immunoreactivity, in the central nervous system of the rat and human Proc. Natl. Acad. Sci. U.S.A. 97, 3655 3660
      88. 88
        Huang, S. M., Bisogno, T., Trevisani, M., Al-Hayani, A., De Petrocellis, L., Fezza, F., Tognetto, M., Petros, T. J., Krey, J. F., Chu, C. J., Miller, J. D., Davies, S. N., Geppetti, P., Walker, J. M., and Di Marzo, V. (2002) An endogenous capsaicin-like substance with high potency at recombinant and native vanilloid VR1 receptors Proc. Natl. Acad. Sci. U.S.A. 99, 8400 8405
      89. 89
        Mishra, S. K., Tisel, S. M., Orestes, P., Bhangoo, S. K., and Hoon, M. A. (2011) TRPV1-lineage neurons are required for thermal sensation EMBO J. 30, 582 593
      90. 90
        Cavanaugh, D. J., Chesler, A. T., Braz, J. M., Shah, N. M., Julius, D., and Basbaum, A. I. (2011) Restriction of transient receptor potential vanilloid-1 to the peptidergic subset of primary afferent neurons follows its developmental downregulation in nonpeptidergic neurons J. Neurosci. 31, 10119 10127
      91. 91
        Lundberg, J. M., Martling, C. R., and Saria, A. (1983) Substance P and capsaicin-induced contraction of human bronchi Acta Physiol. Scand. 119, 49 53
      92. 92
        Mitchell, J. A., Williams, F. M., Williams, T. J., and Larkin, S. W. (1997) Role of nitric oxide in the dilator actions of capsaicin-sensitive nerves in the rabbit coronary circulation Neuropeptides 31, 333 338
      93. 93
        Zygmunt, P. M., Petersson, J., Andersson, D. A., Chuang, H., Sorgard, M., Di Marzo, V., Julius, D., Hogestatt, E. D., and Wang, Y. (1999) Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide Nature 400, 452 457
      94. 94
        Birder, L. A., Nakamura, Y., Kiss, S., Nealen, M. L., Barrick, S., Kanai, A. J., Wang, E., Ruiz, G., De Groat, W. C., Apodaca, G., Watkins, S., and Caterina, M. J. (2002) Altered urinary bladder function in mice lacking the vanilloid receptor TRPV1 Nat. Neurosci. 5, 856 860
      95. 95
        Cruz, F. and Dinis, P. (2007) Resiniferatoxin and botulinum toxin type A for treatment of lower urinary tract symptoms Neurourol. Urodyn. 26, 920 927
      96. 96
        Varga, A., Nemeth, J., Szabo, A., McDougall, J. J., Zhang, C., Elekes, K., Pinter, E., Szolcsanyi, J., and Helyes, Z. (2005) Effects of the novel TRPV1 receptor antagonist SB366791 in vitro and in vivo in the rat Neurosci. Lett. 385, 137 142
      97. 97
        Gavva, N. R. (2008) Body-temperature maintenance as the predominant function of the vanilloid receptor TRPV1 Trends Pharmacol. Sci. 29, 550 557
      98. 98
        Lehto, S. G., Tamir, R., Deng, H., Klionsky, L., Kuang, R., Le, A., Lee, D., Louis, J. C., Magal, E., Manning, B. H., Rubino, J., Surapaneni, S., Tamayo, N., Wang, T., Wang, J., Wang, J., Wang, W., Youngblood, B., Zhang, M., Zhu, D., Norman, M. H., and Gavva, N. R. (2008) Antihyperalgesic effects of (R,E)-N-(2-hydroxy-2,3-dihydro-1H-inden-4-yl)-3-(2-(piperidin-1-yl)-4-(trifluorom ethyl)phenyl)-acrylamide (AMG8562), a novel transient receptor potential vanilloid type 1 modulator that does not cause hyperthermia in rats J. Pharmacol. Exp. Ther. 326, 218 229
      99. 99
        Pabbidi, R. M., Yu, S. Q., Peng, S., Khardori, R., Pauza, M. E., and Premkumar, L. S. (2008) Influence of TRPV1 on diabetes-induced alterations in thermal pain sensitivity Mol. Pain 4, 9
      100. 100
        Van Buren, J. J., Bhat, S., Rotello, R., Pauza, M. E., and Premkumar, L. S. (2005) Sensitization and translocation of TRPV1 by insulin and IGF-I Mol. Pain 1, 17
      101. 101
        Bishnoi, M., Bosgraaf, C. A., Abooj, M., Zhong, L., and Premkumar, L. S. (2011) Streptozotocin-induced early thermal hyperalgesia is independent of glycemic state of rats: role of transient receptor potential vanilloid 1(TRPV1) and inflammatory mediators Mol. Pain 7, 52
      102. 102
        Akiba, Y., Kato, S., Katsube, K., Nakamura, M., Takeuchi, K., Ishii, H., and Hibi, T. (2004) Transient receptor potential vanilloid subfamily 1 expressed in pancreatic islet beta cells modulates insulin secretion in rats Biochem. Biophys. Res. Commun. 321, 219 225
      103. 103
        Razavi, R., Chan, Y., Afifiyan, F. N., Liu, X. J., Wan, X., Yantha, J., Tsui, H., Tang, L., Tsai, S., Santamaria, P., Driver, J. P., Serreze, D., Salter, M. W., and Dosch, H. M. (2006) TRPV1+ sensory neurons control beta cell stress and islet inflammation in autoimmune diabetes Cell 127, 1123 1135
      104. 104
        Wang, X., Miyares, R. L., and Ahern, G. P. (2005) Oleoylethanolamide excites vagal sensory neurones, induces visceral pain and reduces short-term food intake in mice via capsaicin receptor TRPV1 J. Physiol. 564, 541 547
      105. 105
        Ahern, G. P. (2013) Transient receptor potential channels and energy homeostasis Trends Endocrinol. Metab. 24, 554 560
      106. 106
        Kindt, S., Vos, R., Blondeau, K., and Tack, J. (2009) Influence of intra-oesophageal capsaicin instillation on heartburn induction and oesophageal sensitivity in man Neurogastroenterol. Motil. 21, 1032 e1082
      107. 107
        Okumi, H., Tashima, K., Matsumoto, K., Namiki, T., Terasawa, K., and Horie, S. (2012) Dietary agonists of TRPV1 inhibit gastric acid secretion in mice Planta Med. 78, 1801 1806
      108. 108
        Raisinghani, M., Pabbidi, R. M., and Premkumar, L. S. (2005) Activation of transient receptor potential vanilloid 1 (TRPV1) by resiniferatoxin J. Physiol. 567, 771 786
      109. 109
        Roberts, J. C., Davis, J. B., and Benham, C. D. (2004) [3H]Resiniferatoxin autoradiography in the CNS of wild-type and TRPV1 null mice defines TRPV1 (VR-1) protein distribution Brain Res. 995, 176 183
      110. 110
        Chou, M. Z., Mtui, T., Gao, Y. D., Kohler, M., and Middleton, R. E. (2004) Resiniferatoxin binds to the capsaicin receptor (TRPV1) near the extracellular side of the S4 transmembrane domain Biochemistry 43, 2501 2511
      111. 111
        Jung, J., Lee, S. Y., Hwang, S. W., Cho, H., Shin, J., Kang, Y. S., Kim, S., and Oh, U. (2002) Agonist recognition sites in the cytosolic tails of vanilloid receptor 1 J. Biol. Chem. 277, 44448 44454
      112. 112
        Gavva, N. R., Klionsky, L., Qu, Y., Shi, L., Tamir, R., Edenson, S., Zhang, T. J., Viswanadhan, V. N., Toth, A., Pearce, L. V., Vanderah, T. W., Porreca, F., Blumberg, P. M., Lile, J., Sun, Y., Wild, K., Louis, J. C., and Treanor, J. J. (2004) Molecular determinants of vanilloid sensitivity in TRPV1 J. Biol. Chem. 279, 20283 20295
      113. 113
        Brown, D. C., Iadarola, M. J., Perkowski, S. Z., Erin, H., Shofer, F., Laszlo, K. J., Olah, Z., and Mannes, A. J. (2005) Physiologic and antinociceptive effects of intrathecal resiniferatoxin in a canine bone cancer model Anesthesiology 103, 1052 1059
      114. 114
        Iadarola, M. J. and Mannes, A. J. (2011) The vanilloid agonist resiniferatoxin for interventional-based pain control Curr. Top. Med. Chem. 11, 2171 2179
      115. 115
        Yang, B. H., Piao, Z. G., Kim, Y. B., Lee, C. H., Lee, J. K., Park, K., Kim, J. S., and Oh, S. B. (2003) (2003) Activation of vanilloid receptor 1 (VR1) by eugenol J. Dent. Res. 82, 781 785
      116. 116
        Behrendt, H. J., Germann, T., Gillen, C., Hatt, H., and Jostock, R. (2004) Characterization of the mouse cold-menthol receptor TRPM8 and vanilloid receptor type-1 VR1 using a fluorometric imaging plate reader (FLIPR) assay Br. J. Pharmacol. 141, 737 45
      117. 117
        McNamara, F. N., Randall, A., and Gunthorpe, M. J. (2005) Effects of piperine, the pungent component of black pepper, at the human vanilloid receptor (TRPV1) Br. J. Pharmacol. 144, 781 790
      118. 118
        Bhardwaj, R. K., Glaeser, H., Becquemont, L., Klotz, U., Gupta, S. K., and Fromm, M. F. (2002) Piperine, a major constituent of black pepper, inhibits human P-glycoprotein and CYP3A4 J. Pharmacol. Exp. Ther. 302, 645 650
      119. 119
        Bisogno, T., Hanus, L., De Petrocellis, L., Tchilibon, S., Ponde, D. E., Brandi, I., Moriello, A. S., Davis, J. B., Mechoulam, R., and Di Marzo, V. (2001) Molecular targets for cannabidiol and its synthetic analogues: effect on vanilloid VR1 receptors and on the cellular uptake and enzymatic hydrolysis of anandamide Br. J. Pharmacol. 134, 845 852
      120. 120
        Xu, H., Blair, N. T., and Clapham, D. E. (2005) Camphor activates and strongly desensitizes the transient receptor potential vanilloid subtype 1 channel in a vanilloid-independent mechanism J. Neurosci. 25, 8924 8937
      121. 121
        Pearce, L. V., Petukhov, P. A., Szabo, T., Kedei, N., Bizik, F., Kozikowski, A. P., and Blumberg, P. M. (2004) Evodiamine functions as an agonist for the vanilloid receptor TRPV1 Org. Biomol. Chem. 2, 2281 2286
      122. 122
        Morera, E., De Petrocellis, L., Morera, L., Moriello, A. S., Nalli, M., Di Marzo, V., and Ortar, G. (2012) Synthesis and biological evaluation of [6]-gingerol analogues as transient receptor potential channel TRPV1 and TRPA1 modulators Bioorg. Med. Chem. Lett. 22, 1674 1677
      123. 123
        Xu, H., Delling, M., Jun, J. C., and Clapham, D. E. (2006) Oregano, thyme and clove-derived flavors and skin sensitizers activate specific TRP channels Nat. Neurosci. 9, 628 635
      124. 124
        Andre, E., Campi, B., Trevisani, M., Ferreira, J., Malheiros, A., Yunes, R. A., Calixto, J. B., and Geppetti, P. (2006) Pharmacological characterisation of the plant sesquiterpenes polygodial and drimanial as vanilloid receptor agonists Biochem. Pharmacol. 71, 1248 1254
      125. 125
        Iwasaki, Y., Tanabe, M., Kayama, Y., Abe, M., Kashio, M., Koizumi, K., Okumura, Y., Morimitsu, Y., Tominaga, M., Ozawa, Y., and Watanabe, T. (2009) Miogadial and miogatrial with alpha,beta-unsaturated 1,4-dialdehyde moieties—novel and potent TRPA1 agonists Life Sci. 85, 60 69
      126. 126
        Lübbert, M., Kyereme, J., Schöbel, N., Beltrán, L., Wetzel, C. H., and Hatt, H. (2013) Transient receptor potential channels encode volatile chemicals sensed by rat trigeminal ganglion neurons PLoS One 8, e77998
      127. 127
        Trevisan, G., Rossato, M. F., Walker, C. I., Klafke, J. Z., Rosa, F., Oliveira, S. M., Tonello, R., Guerra, G. P., Boligon, A. A., Zanon, R. B., Athayde, M. L., and Ferreira, J. (2012) Identification of the plant steroid alpha-spinasterol as a novel transient receptor potential vanilloid 1 antagonist with antinociceptive properties J. Pharmacol. Exp. Ther. 343, 258 269
      128. 128
        Birnbaumer, L., Yildirim, E., and Abramowitz, J. (2003) A comparison of the genes coding for canonical TRP channels and their M, V and P relatives Cell Calcium 33, 419 432
      129. 129
        Peier, A. M., Reeve, A. J., Andersson, D. A., Moqrich, A., Earley, T. J., Hergarden, A. C., Story, G. M., Colley, S., Hogenesch, J. B., McIntyre, P., Bevan, S., and Patapoutian, A. (2002) A heat-sensitive TRP channel expressed in keratinocytes Science 296, 2046 2049
      130. 130
        Moqrich, A., Hwang, S. W., Earley, T. J., Petrus, M. J., Murray, A. N., Spencer, K. S., Andahazy, M., Story, G. M., and Patapoutian, A. (2005) Impaired thermosensation in mice lacking TRPV3, a heat and camphor sensor in the skin Science 307, 1468 1472
      131. 131
        Xu, H., Ramsey, I. S., Kotecha, S. A., Moran, M. M., Chong, J. A., Lawson, D., Ge, P., Lilly, J., Silos-Santiago, I., Xie, Y., DiStefano, P. S., Curtis, R., and Clapham, D. E. (2002) TRPV3 is a calcium-permeable temperature-sensitive cation channel Nature 418, 181 186
      132. 132
        Smith, G. D., Gunthorpe, M. J., Kelsell, R. E., Hayes, P. D., Reilly, P., Facer, P., Wright, J. E., Jerman, J. C., Walhin, J. P., Ooi, L., Egerton, J., Charles, K. J., Smart, D., Randall, A. D., Anand, P., and Davis, J. B. (2002) TRPV3 is a temperature-sensitive vanilloid receptor-like protein Nature 418, 186 190
      133. 133
        Gopinath, P., Wan, E., Holdcroft, A., Facer, P., Davis, J. B., Smith, G. D., Bountra, C., and Anand, P. (2005) Increased capsaicin receptor TRPV1 in skin nerve fibres and related vanilloid receptors TRPV3 and TRPV4 in keratinocytes in human breast pain BMC Women’s Health 5, 2
      134. 134
        Facer, P., Casula, M. A., Smith, G. D., Benham, C. D., Chessell, I. P., Bountra, C., Sinisi, M., Birch, R., and Anand, P. (2007) Differential expression of the capsaicin receptor TRPV1 and related novel receptors TRPV3, TRPV4 and TRPM8 in normal human tissues and changes in traumatic and diabetic neuropathy BMC Neurol. 7, 11
      135. 135
        Huang, S. M., Lee, H., Chung, M. K., Park, U., Yu, Y. Y., Bradshaw, H. B., Coulombe, P. A., Walker, J. M., and Caterina, M. J. (2008) Overexpressed transient receptor potential vanilloid 3 ion channels in skin keratinocytes modulate pain sensitivity via prostaglandin E2 J. Neurosci. 28, 13727 13737
      136. 136
        Hammarstrom, S., Hamberg, M., Samuelsson, B., Duell, E. A., Stawiski, M., and Voorhees, J. J. (1975) Increased concentrations of nonesterified arachidonic acid, 12L-hydroxy-5,8,10,14-eicosatetraenoic acid, prostaglandin E2, and prostaglandin F2alpha in epidermis of psoriasis Proc. Natl. Acad. Sci. U.S.A. 72, 5130 5134
      137. 137
        Brash, A. R. (2001) Arachidonic acid as a bioactive molecule J. Clin. Invest. 107, 1339 1345
      138. 138
        Mandadi, S., Sokabe, T., Shibasaki, K., Katanosaka, K., Mizuno, A., Moqrich, A., Patapoutian, A., Fukumi-Tominaga, T., Mizumura, K., and Tominaga, M. (2009) TRPV3 in keratinocytes transmits temperature information to sensory neurons via ATP Pfluegers Arch. 458, 1093 2002
      139. 139
        Xiao, R., Tang, J., Wang, C., Colton, C. K., Tian, J., and Zhu, M. X. (2008) Calcium plays a central role in the sensitization of TRPV3 channel to repetitive stimulations J. Biol. Chem. 283, 6162 6174
      140. 140
        Imura, K. (2007) Influence of TRPV3 mutation on hair growth cycle in mice Biochem. Biophys. Res. Commun. 363, 479 483
      141. 141
        Asakawa, M., Yoshioka, T., Matsutani, T., Hikita, I., Suzuki, M., Oshima, I., Tsukahara, K., Arimura, A., Horikawa, T., Hirasawa, T., and Sakata, T. (2006) Association of a mutation in TRPV3 with defective hair growth in rodents J. Invest. Dermatol. 126, 2664 2672
      142. 142
        Macpherson, L. J., Hwang, S. W., Miyamoto, T., Dubin, A. E., Patapoutian, A., and Story, G. M. (2006) More than cool: promiscuous relationships of menthol and other sensory compounds Mol. Cell Neurosci. 32, 335 343
      143. 143
        Moussaieff, A., Rimmerman, N., Bregman, T., Straiker, A., Felder, C. C., Shoham, S., Kashman, Y., Huang, S. M., Lee, H., Shohami, E., Mackie, K., Caterina, M. J., Walker, J. M., Fride, E., and Mechoulam, R. (2008) Incensole acetate, an incense component, elicits psychoactivity by activating TRPV3 channels in the brain FASEB J. 22, 3024 3034
      144. 144
        De Petrocellis, L., Orlando, P., Moriello, A. S., Aviello, G., Stott, C., Izzo, A. A., and Di Marzo, V. (2012) Cannabinoid actions at TRPV channels: effects on TRPV3 and TRPV4 and their potential relevance to gastrointestinal inflammation Acta Physiol. 204, 255 266
      145. 145
        Strotmann, R., Harteneck, C., Nunnenmacher, K., Schultz, G., and Plant, T. D. (2000) OTRPC4, a nonselective cation channel that confers sensitivity to extracellular osmolarity Nat. Cell Biol. 2, 695 702
      146. 146
        Watanabe, H., Davis, J. B., Smart, D., Jerman, J. C., Smith, G. D., Hayes, P., Vriens, J., Cairns, W., Wissenbach, U., Prenen, J., Flockerzi, V., Droogmans, G., Benham, C. D., and Nilius, B. (2002) Activation of TRPV4 channels (hVRL-2/mTRP12) by phorbol derivatives J. Biol. Chem. 277, 13569 13577
      147. 147
        Suzuki, M., Mizuno, A., Kodaira, K., and Imai, M. (2003) Impaired pressure sensation in mice lacking TRPV4 J. Biol. Chem. 278, 22664 22668
      148. 148
        Nilius, B., Droogmans, G., and Wondergem, R. (2003) Transient receptor potential channels in endothelium: solving the calcium entry puzzle? Endothelium 10, 5 15
      149. 149
        Kohler, R., Heyken, W. T., Heinau, P., Schubert, R., Si, H., Kacik, M., Busch, C., Grgic, I., Maier, T., and Hoyer, J. (2006) Evidence for a functional role of endothelial transient receptor potential V4 in shear stress-induced vasodilatation Arterioscler., Thromb., Vasc. Biol. 26, 1495 15502
      150. 150
        Watanabe, H., Vriens, J., Suh, S. H., Benham, C. D., Droogmans, G., and Nilius, B. (2002) Heat-evoked activation of TRPV4 channels in a HEK293 cell expression system and in native mouse aorta endothelial cells J. Biol. Chem. 277, 47044 47051
      151. 151
        Watanabe, H., Vriens, J., Prenen, J., Droogmans, G., Voets, T., and Nilius, B. (2003) Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels Nature 424, 434 438
      152. 152
        Loukin, S. H., Su, Z., and Kung, C. (2009) Hypotonic shocks activate rat TRPV4 in yeast in the absence of polyunsaturated fatty acids FEBS Lett. 583, 754 758
      153. 153
        Cao, D. S., Yu, S. Q., and Premkumar, L. S. (2009) Modulation of transient receptor potential Vanilloid 4-mediated membrane currents and synaptic transmission by protein kinase C Mol. Pain 5, 5
      154. 154
        Matthews, B. D., Thodeti, C. K., Tytell, J. D., Mammoto, A., Overby, D. R., and Ingber, D. E. (2010) Ultra-rapid activation of TRPV4 ion channels by mechanical forces applied to cell surface beta1 integrins Integr. Biol. 2, 435 442
      155. 155
        Alessandri-Haber, N., Dina, O. A., Joseph, E. K., Reichling, D. B., and Levine, J. D. (2008) Interaction of transient receptor potential vanilloid 4, integrin, and SRC tyrosine kinase in mechanical hyperalgesia J. Neurosci. 28, 1046 1057
      156. 156
        Chen, X., Alessandri-Haber, N., and Levine, J. D. (2007) Marked attenuation of inflammatory mediator-induced C-fiber sensitization for mechanical and hypotonic stimuli in TRPV4–/– mice Mol. Pain 3, 31
      157. 157
        Alessandri-Haber, N., Yeh, J. J., Boyd, A. E., Parada, C. A., Chen, X., Reichling, D. B., and Levine, J. D. (2003) Hypotonicity induces TRPV4-mediated nociception in rat Neuron 39, 497 511
      158. 158
        Alessandri-Haber, N., Joseph, E., Dina, O. A., Liedtke, W., and Levine, J. D. (2005) TRPV4 mediates pain-related behavior induced by mild hypertonic stimuli in the presence of inflammatory mediator Pain 118, 70 79
      159. 159
        Smith, P. L., Maloney, K. N., Pothen, R. G., Clardy, J., and Clapham, D. E. (2006) Bisandrographolide from Andrographis paniculata activates TRPV4 channels J. Biol. Chem. 281, 29897 29904

    Pair your accounts.

    Export articles to Mendeley

    Get article recommendations from ACS based on references in your Mendeley library.

    Pair your accounts.

    Export articles to Mendeley

    Get article recommendations from ACS based on references in your Mendeley library.

    You’ve supercharged your research process with ACS and Mendeley!

    STEP 1:
    Click to create an ACS ID

    Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

    Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

    Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

    MENDELEY PAIRING EXPIRED
    Your Mendeley pairing has expired. Please reconnect